You are on page 1of 298

V. 115, NO.

1
JANUARY 2018

ACI
STRUCTURAL J O U R N A L

A JOURNAL OF THE AMERICAN CONCRETE INSTITUTE


CONTENTS
STATEMENT OF OWNERSHIP
Board of Direction ACI Structural Journal
President
Khaled Awad January 2018, V. 115, No. 1
Vice Presidents a journal of the american concrete institute
David A. Lange an international technical society
Randall W. Poston

Directors 3 Editors-in-Chief of ACI Journals Named


JoAnn P. Browning
Cesar A. Constantino 4 Letter from the Editor-in-Chief
Frances T. Griffith
H. R. Trey Hamilton
R. Doug Hooton 5 Effect of Geometric Scaling on Shear Strength of Reinforced
Joe Hug Concrete Beams without Stirrups, by Derek Daluga, Kaylor McCain,
Kimberly Kayler Matthew Murray, and Santiago Pujol
William M. Klorman
Neven Krstulovic-Opara 15 Three-Dimensional Grid Strut-and-Tie Model Approach in Struc-
Tracy D. Marcotte tural Concrete Design, by Young Mook Yun, Byunghun Kim, and
Antonio Nanni Julio A. Ramirez
Roberto Stark

Past President Board Members 27 Verification of Three-Dimensional Grid Strut-and-Tie Model Approach
Michael J. Schneider in Structural Concrete, by Young Mook Yun, Hyun Soo Chae, Byunghun
Sharon L. Wood Kim, and Julio A. Ramirez
William E. Rushing Jr.
41 Time-Dependent Buckling Testing of Eccentrically Loaded Slender
Executive Vice President High-Strength Concrete Panels, by Yue Huang, Ehab Hamed, Zhen-
Ron Burg Tian Chang, and Stephen J. Foster
Editorial Board 53 Nonlinear Backbone Modeling of Concrete Columns Retrofitted
Robert J. Frosch, Editor-in-Chief
  Purdue University with Fiber-Reinforced Polymer or Steel Jackets, by José C. Alvarez,
Catherine French Sergio F. Breña, and Sanjay R. Arwade
  University of Minnesota
Michael Kreger 65 
Shear-Friction Strength of Low-Rise Walls with 550 MPa
  University of Alabama (80 ksi) Reinforcing Bars under Cyclic Loading, by Jang-Woon Baek,
David Sanders Hong-Gun Park, Byung-Soo Lee, and Hyun-Mock Shin
  University of Nevada, Reno
James Wight
  University of Michigan 79 Behavior of Straight and T-Headed ASTM A1035/A1035M Bar
Splices in Flexural Members, by Sergio F. Breña, Jeffrey Messier, and
Staff Sean W. Peterfreund
Publisher
John C. Glumb 91 End-Region Behavior of Pretensioned I-Girders Employing 0.7 in.
(17.8 mm) Strands, by J. Salazar, H. Yousefpour, R. Alirezaei Abyaneh,
Engineering H. Kim, A. Katz, T. Hrynyk, and O. Bayrak
Managing Director
Michael L. Tholen
103 Stepwise Bond Model Including Unconfined and Partially Confined
Managing Editor Hooks, by Armin Erfanian and Alaa E. Elwi
Jerzy Z. Zemajtis
113 Seismic Performance of Innovative Reinforced Concrete Coupling
Staff Engineers
Katie A. Amelio
Beam—Double-Beam Coupling Beam, by Youngjae Choi, Poorya Hajy-
Robert M. Howell alikhani, and Shih-Ho Chao
Khaled Nahlawi
Marc M. Rached 127 A New Approach to Modeling Tension Stiffening in Reinforced
Matthew R. Senecal Concrete, by Angus Murray, Raymond Ian Gilbert, and Arnaud Castel
Gregory M. Zeisler
Publishing Services
Manager
Barry M. Bergin Contents cont. on next page

Editors
Carl R. Bischof Discussion is welcomed for all materials published in this issue and will appear ten months from
this journal’s date if the discussion is received within four months of the paper’s print publication.
Kaitlyn J. Dobberteen Discussion of material received after specified dates will be considered individually for publication or
Tiesha Elam private response. ACI Standards published in ACI Journals for public comment have discussion due
Kelli R. Slayden dates printed with the Standard.
Editorial Assistant ACI Structural Journal
Angela R. Matthews Copyright © 2018 American Concrete Institute. Printed in the United States of America.
The ACI Structural Journal (ISSN 0889-3241) is published bimonthly by the American Concrete Institute. Publication
office: 38800 Country Club Drive, Farmington Hills, MI 48331. Periodicals postage paid at Farmington, MI, and at
additional mailing offices. Subscription rates: $172 per year (U.S. and possessions), $181 (elsewhere), payable in
advance. POSTMASTER: Send address changes to: ACI Structural Journal, 38800 Country Club Drive, Farmington
Hills, MI 48331.
Canadian GST: R 1226213149.
Direct correspondence to 38800 Country Club Drive, Farmington Hills, MI 48331. Telephone: +1.248.848.3700.
Facsimile (FAX): +1.248.848.3701. Website: http://www.concrete.org.

ACI Structural Journal/January 20181


Contributions to
CONTENTS ACI Structural Journal
The ACI Structural Journal is an open
forum on concrete technology and papers
139 Punching Shear Strength of Slabs and Influence of Low Reinforcement related to this field are always welcome.
Ratio, by Susanto Teng, Khatthanam Chanthabouala, Darren T. Y. Lim, and All material submitted for possible publi-
Rhahmadatul Hidayat cation must meet the requirements of
the “American Concrete Institute Publi-
cation Policy” and “Author Guidelines
151 Biaxial Interaction and Load Contour Method of Reinforced Concrete and Submission Procedures.” Prospective
T-Shaped Structural Walls, by Tae-Sung Eom, Hye-Sung Nam, and authors should request a copy of the Policy
Su-Min Kang and Guidelines from ACI or visit ACI’s
website at www.concrete.org prior to
163 Eccentric Punching Shear of Waffle Slab, by Ahmed Faleh Al-Bayati, submitting contributions.
Lau Teck Leong, and Leslie A. Clark Papers reporting research must include
a statement indicating the significance of
175 Modeling Time-Dependent Deformations: Application for Reinforced the research.
Concrete Beams with Recycled Concrete Aggregates, by Adam M.
The Institute reserves the right to return,
Knaack and Yahya C. Kurama
without review, contributions not meeting
the requirements of the Publication Policy.
191 Carbon Fiber-Reinforced Polymer-Strengthened Reinforced Concrete
Beams Subjected to Differential Settlement, by Yail J. Kim and All materials conforming to the Policy
Aiham Al-Kubaisi requirements will be reviewed for editorial
quality and technical content, and every
effort will be made to put all acceptable
203 Tests of Short Headed Bars with Anchor Reinforcement Used in Beam- papers into the information channel.
to-Column Joints, by Ján Bujn 
ák and Matúš Farbák However, potentially good papers may be
returned to authors when it is not possible
211 Evaluation of Flexural and Shear Stiffness of Concrete Squat Walls to publish them in a reasonable time.
Reinforced with Glass Fiber-Reinforced Polymer Bars, by Ahmed Arafa,
Ahmed Sabry Farghaly, and Brahim Benmokrane Discussion
All technical material appearing in the
223 Long-Term Multipliers and Deformability of Fiber-Reinforced Polymer ACI Structural Journal may be discussed.
Prestressed Concrete, by Yail J. Kim and Raymon W. Nickle If the deadline indicated on the contents
page is observed, discussion can appear
235 Seismic Performance of Reinforced Concrete Columns with Lap in the designated issue. Discussion should
Splices in Plastic Hinge Region, by Chul-Goo Kim, Hong-Gun Park, and be complete and ready for publication,
Tae-Sung Eom including finished, reproducible illustra-
tions. Discussion must be confined to the
247 Conventional and High-Strength Steel Hooked Bars: Detailing Effects, scope of the paper and meet the ACI Publi-
by J. Sperry, D. Darwin, M. O’Reilly, A. Lepage, R. D. Lequesne, A. Matam- cation Policy.
oros, L. R. Feldman, S. Yasso, N. Searle, M. DeRubeis, and A. Ajaam Follow the style of the current issue.
Be brief—1800 words of double spaced,
259 
Thermomechanical Relaxation of Reinforced Concrete Beams typewritten copy, including illustrations
Strengthened with Carbon Fiber-Reinforced Polymer, by Yail J. Kim and and tables, is maximum. Count illustrations
Abdulaziz Alqurashi and tables as 300 words each and submit
them on individual sheets. As an approxi-
269 Practicability of Large-Scale Reinforced Concrete Beams Using Grade mation, 1 page of text is about 300 words.
Submit one original typescript on 8-1/2 x
80 Stirrups, by Jung-Yoon Lee, Jae-Hoon Lee, Do Hyung Lee, Seong-Jun
11 plain white paper, use 1 in. margins,
Hong, and Ho-Young Kim and include two good quality copies of the
entire discussion. References should be
281 Use of Anchored Carbon Fiber-Reinforced Polymer Strips for Shear complete. Do not repeat references cited
Strengthening of Large Girders, by William A. Shekarchi, Wassim M. in original paper; cite them by original
Ghannoum, and James O. Jirsa number. Closures responding to a single
discussion should not exceed 1800-word
equivalents in length, and to multiple
discussions, approximately one half of
the combined lengths of all discussions.
Closures are published together with
the discussions.
ON COVER: 115-S22, p. 271, Fig. 1—Test setup and measurement. Discuss the paper, not some new or
outside work on the same subject. Use
references wherever possible instead of
repeating available information.
Discussion offered for publication should
offer some benefit to the general reader.
Discussion which does not meet this
Permission is granted by the American Concrete Institute for libraries and other users registered with the Copyright requirement will be returned or referred to
Clearance Center (CCC) to photocopy any article contained herein for a fee of $3.00 per copy of the article. Payments
should be sent directly to the Copyright Clearance Center, 21 Congress Street, Salem, MA 01970. ISSN 0889-3241/98 the author for private reply.
$3.00. Copying done for other than personal or internal reference use without the express written permission of the
American Concrete Institute is prohibited. Requests for special permission or bulk copying should be addressed to the
Managing Editor, ACI Structural Journal, American Concrete Institute. Send manuscripts to:
The Institute is not responsible for statements or opinions expressed in its publications. Institute publications are not able http://mc.manuscriptcentral.com/aci
to, nor intend to, supplant individual training, responsibility, or judgment of the user, or the supplier, of the information
presented.
Send discussions to:
Papers appearing in the ACI Structural Journal are reviewed according to the Institute’s Publication Policy by individuals
expert in the subject area of the papers. Journals.Manuscripts@concrete.org

2 ACI Structural Journal/January 2018


EDITORS-IN-CHIEF OF ACI JOURNALS NAMED
ACI has formed Editorial Boards to oversee the ACI concrete construction worldwide,” said Khaled W. Awad,
Materials Journal and the ACI Structural Journal. Each ACI President. “By selecting and publishing the most
Editorial Board will comprise an editor-in-chief and four current research with the greatest impact on concrete design
research expert members. The focus of the Editorial Boards and construction, ACI Editorial Boards will provide access
will be on making and keeping ACI’s serial publications the to professionals around the world to the latest information
preferred publishing venue for academic researchers in the on concrete and its uses. I would like to take the opportunity
field of concrete. The Boards’ initial missions will include to thank the world-renowned experts who accepted to serve
identifying new topics for commissions and special editions, on the Boards for their passion and dedication to support
advising on the direction for the journals, making sugges- ACI in advancing concrete knowledge globally.”
tions for both subject matter and potential authors, providing The ACI Materials Journal and ACI Structural Journal
content by writing occasional editorials and other short arti- have been publishing research on properties of concrete
cles, and overseeing journal quality. materials and design of structural concrete since ACI began
Jason Weiss, Oregon State University, serves as Editor- publishing journals back in December 1912. The goal has
in-Chief of the ACI Materials Journal Editorial Board, always been to ensure that those who design and build
with Board members Zachary Grasley, Texas A&M Univer- with concrete have access to the most current and cutting-
sity; Maria Juenger, University of Texas at Austin; Kamal edge research and information on concrete. In many cases,
Khayat, Missouri S&T; and Michael Thomas, University of research published in ACI’s journals directly impacts the
New Brunswick. codes, specifications, and practices published by the Insti-
Robert Frosch, Purdue University, serves as Editor-in- tute and other standards developing organizations.
Chief of the ACI Structural Journal Editorial Board, with To date, more than 4700 papers have been published in the
Board members Catherine French, University of Minnesota; print and digital editions of the ACI Materials Journal and
Michael Kreger, University of Alabama; David Sanders, ACI Structural Journal. Subscriptions to the digital editions
University of Nevada, Reno; and James Wight, University of both journals are included with ACI membership.
of Michigan.
“The formation of the Editorial Boards for the ACI Reprinted from the November 2017 issue of Concrete
Materials Journal and ACI Structural Journal will further International, V. 39, No. 11, p. 12.
support the advancement of concrete technology and

ACI Structural Journal/January 2018 3


LETTER FROM THE EDITOR-IN-CHIEF
This issue (Volume 115, No. 1) marks a new era in the evolu- publish for academic researchers in the field of concrete. The
tion of the ACI Structural Journal. ACI began publishing in Boards’ initial missions include identifying new topics for
1905 with the Proceedings of the First Convention of the commissions; special editions; and advising on the direction
National Association of Cement Users, which was held in for the journals, making suggestions for both subject matter
Indianapolis, IN. For the next century, ACI has continued and potential authors, providing content by writing occa-
to advance the profession through dissemination of concrete sional editorials and other short articles, overseeing journal
knowledge and research. This is no small feat and shows the quality, and more.
dedication of ACI to advance technical knowledge in the field I am very excited about the opportunities that this new
of concrete. Having such a long history is quite remarkable, management structure will provide. This is the first time in
considering the changes that have occurred in technology the history of the Institute that an Editorial Board has been
and society over the 115 volumes of the Journal. To date, formed. As Editor-in-Chief, I plan to explore ways that
ACI has published more than 9000 papers and discussions we can further strengthen and enhance the reputation of
between the ACI Structural Journal, ACI Materials Journal, the ACI Structural Journal true to ACI’s tagline “Always
and the ACI Journal Proceedings (which predated the sepa- Advancing.” It is often said that “change is the constant,”
rate Structural and Materials Journals). These publications and that is no different with scholarly publications. Changes
have influenced building codes, design specifications, and are continually occurring in the means and methods of publi-
construction practice and resulted in a more economic and cation and the manner of dissemination due to electronic and
safer built environment. The goal has always been to ensure social media. The ACI Journals will continue to evolve to
that those who design and build with concrete have access to remain relevant in this ever-changing world. I believe this is
the most current and cutting-edge research and information an exciting time to explore new approaches, allowing us to
on concrete. continue to advance the state of knowledge and improve the
As time has progressed, the Journal has seen numerous built environment.
changes from preparation (manuscripts prepared by type-
writers being submitted through the mail to electronic prepa- Robert J. Frosch, PhD, PE, FACI
ration and submission) to dissemination (hard-copy mailings Editor-in-Chief
to electronic distribution). Management of the Journal has Purdue University
also evolved over the years. Starting with this issue, both
the ACI Structural and Materials Journals will be overseen p.s. A couple of interesting notes:
by an Editorial Board. The Editorial Board is comprised of
an editor-in-chief and four research expert members. As 1. Amazingly, ACI has scanned every publication
Editor-in Chief of the Structural Journal, I am honored to going back to Volume 1 in 1905. ACI members can
be joined with an outstanding Editorial Board that includes download for free any of the more than 9000 papers
Catherine French, Michael Kreger, David Sanders, and and discussions through the ACI International Concrete
James Wight. I am sure that that these individuals are known Abstracts Portal: https://www.concrete.org/publications/
to you based on their excellent research and contributions to internationalconcreteabstractsportal.
both the profession and the Institute. 2. Information regarding submitting manuscripts is
The Editorial Boards will focus on the mission of making covered in detail through the Publications Portal: https://
and keeping ACI’s serial publications the premier place to www.concrete.org/publications/acistructuraljournal.

4 ACI Structural Journal/January 2018


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 114-S113
Note: Paper 114-S113 of the November-December 2017 ACI Structural Journal was printed with the erroneous title: “Effect of
Geometric Scaling on Shear Strength of Reinforced Concrete Beams with Stirrups.” A reprint including the corrected title is included
herein. Please reference the January 2018 version of this paper only.

Effect of Geometric Scaling on Shear Strength of


Reinforced Concrete Beams without Stirrups
by Derek Daluga, Kaylor McCain, Matthew Murray, and Santiago Pujol

The experimental work and studies reported in this article were


undertaken to study the effects of dimensional changes on the unit
shear strength of reinforced concrete beams. The tests included the
following variables: beam depth; maximum aggregate size; and
bar cover, size, and spacing.
Studies of the experimental results suggested that if changes
in the considered variables are controlled so that the variables
change in the same proportion, the variation in the experimental
results is not more than the expected scatter in results from nomi-
nally identical beams.

Keywords: maximum aggregate size; reinforced concrete; scaling; shear;


size effect.

INTRODUCTION
Although the work of ACI-ASCE Committee 326 has
helped the industry with respect to shear and diagonal
tension, there are cases, albeit infrequent, where beams
and slabs without stirrups and depths much larger than the
depths of the beams considered by Committee 326 (depths
exceeding 36 in. [914 mm]) are built, and the reliability of
such elements has been questioned.1-3 There are many alter-
natives to theories explaining size effect4-6; however, in this
paper, the focus was placed on experimental results.
Current design methods related to shear are based predom-
inantly on results from tests on 194 beams with depths
smaller than 24 in. (610 mm).7-9 Eighty-eight percent of these Fig. 1—Shear test results.16
beams had depths less than 16 in. (406 mm). Later experi- The majority of previous investigations on the relation-
ments included beams with larger depths.1,10-15 The results of ship between shear strength and depth focused on beams
these experiments have been interpreted to suggest that, as with similar shear span-to-effective depth and reinforcement
beam depth increases from 12 to 36 in. (305 to 914 mm), the ratios. However, ratios of depth to maximum aggregate size
unit shear strength can decrease by as much as 68%.14 This and depth to reinforcement diameter, spacing, and cover
phenomenon, often termed “size effect”, is more noticeable differed. The ratio of beam depth to width is not mentioned
in beams without shear reinforcement,12 and is illustrated in because previous investigations suggest that beam width does
Fig. 1 using data collected by Reineck et al.16 The data in this not have a perceptible effect on unit shear strength.1,17,18 On
figure have parameters in these ranges the other hand, tests have shown that decreasing aggregate
size has a negative effect on shear capacity, bar diameter is
2.3 ≤ a/d ≤ 4.0 likely to affect dowel force, and crack width and spacing are
3/8 in. (10 mm) ≤ ag ≤ 1.2 in. (30 mm) sensitive to bar cover and spacing.13,19-22 These three issues
are discussed briefly in the next sections.
0.01 ≤ ag/d ≤ 0.2
3400 psi (23 MPa) ≤ fc′ ≤ 10,800 psi (74 MPa) ACI Structural Journal, V. 115, No. 1, January 2018.
MS No. S-2015-318.R2, doi: 10.14359/51700947, received August 16, 2016, and
6 in. (152 mm) ≤ h ≤ 49 in. (1245 mm) reviewed under Institute publication policies. Copyright © 2018, American Concrete
Institute. All rights reserved, including the making of copies unless permission is
obtained from the copyright proprietors. Pertinent discussion including author’s
0.6% ≤ ρ ≤ 2.9% closure, if any, will be published ten months from this journal’s date if the discussion
is received within four months of the paper’s print publication.

ACI Structural Journal/January 2018 5


Aggregate size be pumped to high stories). Aggregate is scaled in this inves-
The sensitivity of shear strength to maximum aggregate tigation for academic purposes.
size is illustrated in Fig. 2 using the same data from Reineck The trend in Fig. 2 also suggests that aggregate size can
et al.16 in Fig. 1. Maximum aggregate size is defined herein affect aggregate interlock, which affects shear strength. If
as the size of the smallest sieve opening through which all variables remain equal, as depth increases, crack width
100% of particles pass. The data illustrated in Fig. 2 suggest increases as well. If aggregate size does not increase at
that, for normalweight aggregate (such that the intrinsic the same rate as depth (and subsequently crack width), its
aggregate strength is capable of transferring shear across the capacity to transfer shear across a crack decreases.
cracks), shear strength increases with an increase in the ratio
of reported maximum aggregate size to depth. This trend Bar diameter
suggests that, to avoid large reductions in shear strength Shear force resisted by longitudinal reinforcement, or
with increases in depth, aggregate size should be increased dowel force, is affected by bar diameter. If dowel force is
in large beams. This is often not feasible in the field. For at least in part caused by bending of the longitudinal rein-
instance, if a mixture with common 3/4 in. (19 mm) aggregate forcement as it spans an inclined crack, a larger number of
is used for a 12 in. (305 mm) deep beam, then a comparable smaller bars with the same total cross-sectional area as a
48 in. (1219 mm) beam would require a 3 in. (76 mm) aggre- smaller number of larger bars would be expected to transfer
gate mixture. To achieve a workable mixture with aggregate less shear across an inclined crack (at least before a splitting
of this size is difficult (especially for mixtures that need to crack forms). In beams of equal size, equal reinforcement
ratio, and different bar size, the number of bars is propor-
tional to the inverse of the area (and the square of bar diam-
eter). The bending stiffness of a single bar is proportional to
the fourth power of its diameter. For the group of bars, this
results in a total lateral stiffness proportional to the square
of bar diameter. For example, if instead of ten No. 4 bars
(with a total area of 2 in.2 [1290 mm2]) one uses two No. 9
bars (resulting in the same area), the bending stiffness of
the group of No. 4 bars is one-fifth (or 20% = 42/92) of the
bending stiffness of the No. 9 bars. In addition, if the bars
are arranged in a single layer, the beam with smaller bars
has less area of concrete resisting tension across the poten-
tial plane of splitting. The lateral stiffness of the longitudinal
bars may also affect crack width, combining the effects of
dowel force and aggregate size.

Bar cover and spacing


To illustrate the relevance of bar cover c and bar spacing s
on cracking, consider tests reported by Sneed and Ramirez.14
They tested beams where concrete cover and bar spacing
were not scaled in proportion to depth. Figure 3 shows a
comparison of crack patterns in beams with different effec-
tive depths d but similar shear span-to-effective depth and
reinforcement ratios.23 These patterns were observed at
loads producing comparable bending stresses at any given
cross section. The observed crack patterns were different,
Fig. 2—Variation of shear strength with relative aggregate size.

Fig. 3—Crack patterns.23

6 ACI Structural Journal/January 2018


with the smaller beam (c/d = 0.3) having fewer cracks than proportion to depth, an increase in depth causes a negligible
the larger beam (c/d = 0.11). Figure 4 shows how the number decrease in unit shear strength. For the reasons presented,
of cracks increased with increases in depth in the tests by ratios of bar spacing, diameter, and cover to depth may be
Sneed.23 It is plausible that differences in crack patterns and important, but in the reported experiments, these ratios were
crack spacing result in differences in relative crack width, kept nearly constant to concentrate on the effects of aggre-
stress distribution, load path, and shear strength. gate size-to-depth ratios.
Taylor24 hypothesized that the ratios of aggregate size and
bar cover, size, and spacing to depth affect shear strength. RESEARCH SIGNIFICANCE
To test this idea, Taylor tested 16 beams in which aggregate Shear strength of beams has been reported to decrease
size, width, and bar cover, size, and spacing were scaled in with increases in depth. This phenomenon is commonly
proportion to depth. All beams had a reinforcement ratio referred to as size effect. From 1900 to 1980, ACI published
of 1.35% and a shear span-to-effective depth ratio of 3.0. four articles on the subject (keyword search: size effect25).
His focus on scaling yielded a 20% decrease in unit shear In the following 35 years, 130 articles related to size effect
strength between beams with depths of 9.4 and 39.4 in. (250 have appeared in ACI publications. More than 20 of these
and 1000 mm). In his test, however, a direct comparison papers were published in the last 5 years, indicating it is an
between beams of different size but similar nondimensional important matter in the profession.
parameters could only be made between six beams. Taylor23 observed that the reported reduction in shear
The purpose of the tests reported in this paper was to test strength decreased dramatically when maximum aggre-
Taylor’s hypothesis that if dimensions including maximum gate size, and bar cover, size, and spacing were scaled in
aggregate size and bar cover, size, and spacing are scaled in proportion to depth. This observation shed new light onto
the problem of size effect. Nevertheless, Taylor’s conclu-
sion seems to have gone mostly unnoticed by the profession,
perhaps because it was based on a small number of tests. The
study reported herein was conducted to test his conclusion,
and to separate the plausible effects of increases in size from
effects related to nondimensional variables.

EXPERIMENTAL PROCEDURE
Thirty reinforced concrete simply supported beams
without shear reinforcement were tested by applying a
concentrated load at midspan until shear failure occurred.
For 13 specimens, external stirrups were used to reinforce
the half of the beam where failure first occurred. Load was
applied after the external stirrups were installed to test to
failure the other half of the beam. Here, results from oppo-
site halves of a test beam are distinguished from one another
by using the suffix -N or -S after the beam ID. The nominal
dimensions and parameters of the beams are listed in Table 1.
The larger beam dimensions were 2.5 and 4 times the dimen-
Fig. 4—Number of flexural cracks.23
Table 1—Beam properties
Effective Area of tension Maximum Bar Length of bearing
Depth h, Width, b, depth, d, in. reinforcement, L, in. s, in. aggregate size diameter db, Cover, c, plates, lb,
Series in. (mm) in. (mm) (mm) As, in.2 (mm2) (mm) (mm) ag, in. (mm) in. (mm) in. (mm) in. (mm)
A 30 (762) 22.5 (572) 26.25 (667) 4.68 (3018) 120 (3048) 7.5 (191) 1 (25) 1.41 (35.8) 3.75 (95) 10 (254)
B 12 (305) 9 (229) 10.5 (267) 0.93 (600) 48 (1219) 3 (76) 1 (25) 0.63 (15.9) 1.50 (38) 4 (102)
C 12 (305) 9 (229) 10.5 (267) 0.60 (387) 48 (1219) 3 (76) 1 (25) 0.5 (12.7) 1.50 (38) 4 (102)
D 12 (305) 9 (229) 10.5 (267) 0.60 (387) 48 (1219) 3 (76) 0.5 (13) 0.5 (12.7) 1.50 (38) 4 (102)
E 30 (762) 22.5 (572) 26.25 (667) 4.68 (3018) 150 (3810) 7.5 (191) 1 (25) 1.41 (35.8) 3.75 (95) 10 (254)
F-1, F-2 12 (305) 9 (229) 10.5 (267) 0.93 (600) 60 (1524) 3 (76) 0.5 (13) 0.63 (15.9) 1.50 (38) 4 (102)
G 48 (1219) 24 (610) 42 (1067) 8.00 (5162) 240 (6096) 12 (305) 2 (51) 2.26 (57.3) 6.00 (152) 10 /12† (254/305)
*

H 12 (305) 9 (229) 10.5 (267) 0.93 (597) 60 (1524) 3 (76) 0.375 (10) 0.63 (15.9) 1.50 (38) 4 (102)
F-3, F-4, I, J 12 (305) 9 (229) 10.5 (267) 0.93 (597) 60 (1524) 3 (76) 0.5 (13) 0.63 (15.9) 1.50 (38) 4 (102)
K 12 (305) 9 (229) 10.5 (267) 0.93 (597) 60 (1524) 3 (76) 1 (25) 0.63 (15.9) 1.50 (38) 4 (102)
*
At support.

At load point.

ACI Structural Journal/January 2018 7


sions of the smallest beams. Table 2 shows the ratios of key diameter-to-depth ratio. Longitudinal reinforcement ratio
beam dimensions (shear span a, bar diameter db, concrete ranged between 0.63% and 0.98%. Cross-sectional dimen-
cover c, reinforcement spacing s, and maximum aggregate sions are shown in Fig. 5. The ratio of width to effective
size ag) to effective depth d. Changes in shear span-to-effec- depth was smaller in the 48 in. (1219 mm) deep beams
tive depth ratio a/d were introduced to study the sensitivity (Series G) so the entire beam could be cast from the same
of results to this parameter. Two different shear span-to-ef- batch. Fewer bars were used in the 48 in. (1219 mm) deep
fective depth ratios were used: 2.3 and 2.9. Other ratios were beams (two instead of three) to keep all nondimensional
nearly constant among all test beams except for beams in ratios within the ranges of comparable beams (Table 2).
Test Series B, C, and K. In Series B, C, and K, the ratio For this reason, the reader is encouraged to compare results
of maximum aggregate size to depth ag/d was larger than in terms of stress (accounting for differences in concrete
in other series (0.095 versus 0.036 to 0.048) to study the strength) instead of force. It is also helpful to realize that the
effects maximum aggregate size. Other differences in ag/d effects of increases in aggregate size and bar size, cover, and
are consequences of limitations in material availability. The spacing, although not always viable in practice, help reduce
same is true for differences in reinforcement ratio and bar differences in aggregate interlock and dowel force (again in
terms of stress, not force).
Table 2—Ratio of beam parameters to effective Testing was conducted at the Robert L. and Terry L. Bowen
depth Laboratory for Large-Scale Civil Engineering Research at
Shear Bar Reinforce- Cover Maximum Purdue University. Figure 6 shows the test setup used for the
span to diameter ment spacing to aggregate 30 and 48 in. (762 and 1219 mm) deep beams. The test setup
depth to depth to depth depth size to depth was similar for the 12 in. (305 mm) deep beams. Rollers
Series a/d db/d s/d c/d ag/d supported the beams at each end of the test span. The sizes
A 2.3 0.054 0.286 0.143 0.038 of plates used to distribute reactions and loads are listed in
Table 3.
B 2.3 0.060 0.286 0.143 0.095
Beams were cast with bars on the bottom. Each individual
C 2.3 0.048 0.286 0.143 0.095 beam was cast with concrete from the same batch. After the
D 2.3 0.048 0.286 0.143 0.048 concrete set, it cured under a layer of burlap and plastic.
E 2.9 0.054 0.286 0.143 0.038 Cure time differed among beams, but all specimen dried for
at least 2 weeks before testing.
F-1, F-2 2.9 0.060 0.286 0.143 0.036
Concrete was made using Type I portland cement and
G 2.9 0.054 0.286 0.143 0.048 aggregate mined in Indiana. Series G was made with crushed
H 2.9 0.048 0.286 0.143 0.036 dolomitic limestone, and the other series were made using
F-3, F-4, I, J 2.9 0.048 0.286 0.143 0.048 gravel from alluvial deposits. High-strength longitudinal
reinforcement (yield stress fy ≥ 100 ksi [690 MPa]) was used
K 2.9 0.048 0.286 0.143 0.095
as needed to prevent yielding before shear failure. Other

Fig. 5—Specimen design and test setup.

8 ACI Structural Journal/January 2018


Table 3—Plate dimensions
Group Bearing plates, in. (mm) Location
B, C, D, F,
9 x 4 x 3/4 (229 x 102 x 19) Supports and load point
H, I, J, K
A, E 22.5 x 10 x 2 (572 x 254 x 51) Supports and load point
24 x 10 x 2 (610 x 254 x 51) Supports
G
36 x 12 x 3 (914 x 305 x 76) Load point

posed in Fig. 8. One was observed in a 12 in. (305 mm)


deep beam and the other in a 48 in. (1219 mm) deep beam.
Both beams had nine to ten flexure cracks at loads producing
similar bending stresses. The number of flexural cracks was
not always the same in the tests reported herein, but on
average, the number of cracks did not change consistently
with changes in depth, as shown in Fig. 9. If there is a trend
in number of cracks, however, it is both difficult to distin-
guish and a much weaker trend than that in Fig. 4. Figure 9
was produced by counting cracks at loads producing similar
bending stresses at load stages close to failure.27
The peak load P measured in the laboratory was used
to calculate nominal peak shear stress V/(bd), both at the
center of the support and at a distance of d (effective depth)
from the center of the support. The calculated shear stress
includes self-weight and weight of the loading equipment
(Table 4). As the specimens had different concrete compres-
sive strengths, unit shear stress is normalized with respect
to the square root of concrete strength (v/√fc′) in Table 4 and
the following comparisons. The square root of fc′ is used to
conform to the conventions adopted in design codes in the
Fig. 6—Loading setup.
United States.28
details, including aggregate gradations and mixture propor- Peak unit shear strength at d from support center for beams
tions, are reported by Murray,9 McCain,26 and Daluga.27 with a shear span-to-effective depth ratio of 2.3 and 2.9 and
Daluga also examined the possibility that tensile concrete ag/d within the narrow range of 0.036 to 0.048 are shown in
strength varied consistently with aggregate size in the Fig. 10(a) and 10(b), respectively. Beams in Series B, C, and
mixtures used. Such variation would have introduced an K (which had a large ratio of aggregate size to depth) are
additional (unintended) variable to the reported test program. excluded from Figure 10 to try to isolate the effects of depth
No clear correlation between tensile concrete strength and increases by keeping other relevant variables within these
aggregate size was detected for the mixtures used. Tensile narrow ranges
concrete strength obtained from split cylinder tests ranged
between 310 and 560 psi (2.1 and 3.9 MPa). 0.63% ≤ ρ ≤ 0.98%
0.036 ≤ ag/d ≤ 0.048
EXPERIMENTAL RESULTS AND DISCUSSION
In each beam, flexural cracks began to form near the s/d = 0.29
midspan. As load increased, additional flexural cracks began c/d = 0.14
to form closer to the supports. These cracks became inclined
toward the load point with increasing loads. All beams failed 2700 psi (19 MPa) ≤ fc′ ≤ 5000 psi (34 MPa)
because of the formation of an inclined crack, resulting in
shear failure. The data in Fig. 10(a) and 10(b) show a reduction in mean
Experimental results are summarized in Table 4. Listed shear strength of 5% and 14%, respectively. This reduction
values of concrete compressive strength fc′ are the means between beams with depths of 12 and 48 in. (305 and 1219 mm)
of the strengths of three 6 x 12 in. (152 x 305 mm) cylin- is smaller than both the spread of the data and reductions
ders tested on the same day as the corresponding beam test. reported in other studies (up to 68%).1,14,29 It is plausible that
Figure 7 shows all the measured load-deflection curves for this reduction in strength is related to differences in casting
beams of different size organized by relative aggregate size and curing conditions between large and small beams. It is
and aspect ratio. also plausible that in larger beams, with larger aggregate
Adequate scaling is deemed to have been achieved particles and reinforcement bars, more bleed water gets
because similar crack patterns were observed in small and trapped under the aggregate particles and bars weakening
larger beams (Fig. 8). Two crack patterns are superim- the concrete around them.

ACI Structural Journal/January 2018 9


Table 4—Summary of experimental program
v, psi* v/√fc′ (fc′ in psi)†
Reinforcement
Specimen fc′, psi ft, psi ratio,% Peak load P, kip At support At d from support At support At d from support
A-1 3600 520 0.79 146 130 127 2.2 2.1
A-2 4700 560 0.79 132 118 116 1.7 1.7
A-3 4000 430 0.79 190 167 165 2.6 2.6
A-4 4500 425 0.79 183 161 159 2.4 2.4
Average 2.2 2.2
B-1 3600 520 0.98 34 183 182 3.0 3.0
B-2 3600 520 0.98 40 215 214 3.6 3.6
Average 3.3 3.3
C-1 4900 560 0.63 44 236 235 3.4 3.4
C-2 4700 560 0.63 39 209 208 3.1 3.0
Average 3.2 3.2
D-1 2700 400 0.63 22 119 118 2.3 2.3
D-2 2800 400 0.63 23 125 124 2.4 2.3
D-3 4400 440 0.63 27 146 145 2.2 2.2
D-4 4500 480 0.63 31 167 166 2.5 2.5
D-5 4500 480 0.63 31 167 166 2.5 2.5
D-6 4600 470 0.63 29 156 155 2.3 2.3
Average 2.4 2.3
E-1 4300 420 0.79 130 118 115 1.8 1.8
E-2 4300 470 0.79 136 123 120 1.9 1.8
E-2-S 4400 470 0.79 142 128 125 1.9 1.9
Average 1.9 1.8
F-1 4900 490 0.98 27 146 145 2.1 2.1
F-2 5000 450 0.98 29 157 156 2.2 2.2
Average 2.2 2.2
F-3-S 4900 510 0.63 27 147 146 2.1 2.1
F-3-N 4900 510 0.63 27 147 146 2.1 2.1
F-4-S 4900 510 0.63 26.6 145 144 2.1 2.1
F-4N 4900 510 0.63 31.2 169 168 2.4 2.4
Average 2.2 2.2
G-1-N 4500 450 0.79 204 114 110 1.7 1.6
G-1-S 4500 450 0.79 233 128 124 1.9 1.9
G-2-N 3400 410 0.79 205 115 110 2.0 1.9
G-2-S 3400 410 0.79 205 115 110 2.0 1.9
Average 1.9 1.8
H-1-N 4100 310 0.63 19.9 110 109 1.7 1.7
H-1-S 4100 310 0.63 24.9 136 135 2.1 2.1
H-2-S 4100 310 0.63 21.6 119 118 1.9 1.8
H-2-N 4100 310 0.63 23.3 128 127 2.0 2.0
Average 1.9 1.9
I-1-S 4400 440 0.63 24.8 136 135 2.0 2.0
I-1-N 4400 440 0.63 30 163 162 2.5 2.4

10 ACI Structural Journal/January 2018


Table 4 (cont.)—Summary of experimental program
I-1-S 4200 440 0.63 23.8 130 129 2.0 2.0
I-1-N 4200 440 0.63 27.2 148 147 2.3 2.3
Average 2.2 2.2
J-1-S 3300 330 0.63 22.9 126 125 2.2 2.2
J-1-N 3300 330 0.63 23.8 130 129 2.3 2.3
J-2-S 3300 330 0.63 22.7 125 123 2.2 2.1
J-2-N 3300 330 0.63 26.9 147 146 2.6 2.5
Average 2.3 2.3
K-1-N 4000 360 0.63 27.5 150 149 2.4 2.4
K-1-S 4000 360 0.63 28.9 157 156 2.5 2.5
K-2-S 4300 450 0.63 29.7 162 161 2.5 2.4
K-2-N 4300 450 0.63 32 174 173 2.6 2.6
Average 2.5 2.5
*
Includes self-weight and weight of loading equipment (500 lb [2.2 kN] for A and E; 100 lb [0.4 kN] for D, F-1, and F-2; 270 lb [1.2 kN] for F-3, F-4, H, I, J, and K; and 2000 lb
[8.9 kN] for G).

To convert this value into terms of MPa, divide it by 12.
Notes: 1 psi = 0.0069 MPa; 1 kip = 4.45 kN.

Fig. 7—Load-deflection curves (x-axis = ∆/L and y-axis = v/√fc′).


To try to isolate the effect of size, the ratio of maximum The effect of aggregate size is illustrated in Fig. 11, which
aggregate size to effective depth was controlled. This is compares results from Table 4 for beams where the only
seldom feasible in the field, but it does help to understand nominal difference is maximum aggregate size. Increases
the causes of the reduction in unit shear strength that has in maximum aggregate size from 3/8 to 1 in. (10 to 25 mm)
often been reported as size effect. As previously explained, yielded increases in shear strength larger than 23% in beams
maximum aggregate size can affect aggregate interlock. with shear span-to-effective depth ratios of 2.3 (Fig. 11(a)) and

ACI Structural Journal/January 2018 11


Fig. 8—Crack patterns.27

Fig. 9—Number of flexural cracks.


2.9 (Fig. 11(b)). This observation suggests that increasing
depth without increasing aggregate size can result in a
reduction in unit shear strength.

CONCLUSIONS
Thirty beams were tested to study the sensitivity of unit
shear strength to increases in cross-sectional depth. Beams
with two shear span-to-effective depth ratios were tested:
a/d = 2.3 and 2.9. In all test series except Series B, C, and
K, maximum aggregate size and bar cover, size, and spacing
were increased in similar proportion to beam depth. Beam
depth ranged from 12 to 48 in. (305 to 1219 mm).
For both shear span-to-effective depth ratios a/d, and
for beams with similar ratios of aggregate size to effective
depth, a reduction in unit shear strength V/bd was observed
with an increase in cross-sectional depth. For beams with
depths of 12 and 48 in. (305 and 1219 mm), mean unit shear
strength decreased approximately 14%. This reduction is
smaller than: 1) the spread of the data; and 2) the reduc-
tion observed in previous studies (up to 68%) (Fig. 12).1,14,29
The smaller reduction in shear strength with depth reported
herein was observed in beams within these ranges:

2.3 ≤ a/d ≤ 2.9


0.63% ≤ ρ ≤ 0.98%
0.036 ≤ ag/d ≤ 0.048
s/d = 0.29 Fig. 10—Variation of unit shear strength with beam depth
(mean unit shear strength marked in solid line).
c/d = 0.14

12 ACI Structural Journal/January 2018


Fig. 12—Comparison of data from Fig. 1 and data from this
investigation.
Matthew Murray is a Project Manager at CE Solutions, Inc., Carmel, IN.
He received his BS and MS in civil engineering from Purdue University
in 2009 and 2010, respectively. His research interests include structural
design of new construction; additions; and repair and restoration of indus-
trial, environmental, commercial, and residential structures.

Santiago Pujol, FACI, is a Professor of civil engineering at Purdue Univer-


sity. He is a member of ACI Committees 133, Disaster Reconnaissance;
314, Simplified Design of Concrete Buildings; and 318, Structural Concrete
Building Code; and Joint ACI-ASCE Committees 441, Reinforced Concrete
Columns; and 445, Shear and Torsion.

ACKNOWLEDGMENTS
The authors express their gratitude to M. Sozen and J. Ramírez for their
help with the project and critical reviews of the results, Bowen Labora-
tory staff (K. Brower and H. Tidrick) and research assistants for their help
with tests, and MMFX for its generous donations. The constructive criti-
cism kindly provided by A. B. Acevedo at EAFIT, Colombia, is gratefully
acknowledged.

Fig. 11—Variation of shear strength with aggregate size


NOTATION
(d = 10.5 in. [267 mm]). As = cross-sectional area of reinforcing bars
a = shear span
2700 psi (19 MPa) ≤ fc′ ≤ 5000 psi (34 MPa) ag = maximum aggregate size
b = cross-sectional width
12 in. (305 mm) ≤ h ≤ 48 in. (1219 mm) c = distance from nearest concrete face to centroid of longitudinal
reinforcement
d = effective depth
The reported observations support those made by Taylor24 db = bar diameter
and suggest that the problem commonly referred to as size fc = shear stress at comparable loading stage (P/bd)
effect is an effect more closely related to scaling, particu- fc′ = compressive strength of standard 6 x 12 in. (152 x 305 mm)
concrete cylinder
larly of the maximum aggregate size, than to size itself. h = cross-sectional height
L = length of beam between centers of supports
AUTHOR BIOS lb = length of bearing plate
ACI member Derek Daluga received his BS and MS in civil engineering m = length of overhanging portion of beam
from Purdue University, West Lafayette, IN, in 2013 and 2015, respectively. P = applied concentrated load
His research interests include the effects of aggregate size on the shear s = longitudinal reinforcement spacing, center-to-center
strength of reinforced concrete. V = shear force
v = unit shear stress (V/bd)
ACI member Kaylor McCain is a Design Engineer at Structural Design ∆ = deflection
Group, Nashville, TN. He received his BS in civil engineering from Auburn ρ = longitudinal reinforcement ratio (As/bd)
University, Auburn, AL, in 2010, and MS in civil engineering from Purdue
University in 2012. His research interests include size effect and the design
and strengthening of reinforced concrete structures.
REFERENCES
1. Kani, G. N. J., “How Safe Are Our Large Reinforced Concrete
Beams?” ACI Journal Proceedings, V. 64, No. 3, Mar. 1967, pp. 128-141.

ACI Structural Journal/January 2018 13


2. Collins, M. P., and Kuchma, D. A., “How Safe Are Our Large, Lightly 16. Reineck, K.; Kuchma, D. A.; Kim, K. S.; and Marx, S., “Shear Data-
Reinforced Concrete Beams, Slabs, and Footings?” ACI Structural Journal, base for Reinforced Concrete Members without Shear Reinforcement,” ACI
V. 96, No. 4, July-Aug. 1999, pp. 482-490. Structural Journal, V. 100, No. 2, Mar.-Apr. 2003, pp. 240-249.
3. Collins, M. P.; Bentz, E. C.; Quach, P. T.; and Proestos, G. T., “The 17. Lubell, A.; Sherwood, E.; Bentz, E. C.; and Collins, M. P., “Safe
Challenges of Predicting the Shear Strength of Very Thick Slabs,” Concrete Shear Design of Large, Wide Beams,” Concrete International, V. 26,
International, V. 37, No. 11, Nov. 2015, pp. 29-37. No. 1, Jan. 2004, pp. 66-78.
4. Bažant, Z. P., and Kim, J., “Size Effect in Shear Failure of Longi- 18. Sherwood, E. G.; Lubell, A.; Bentz, E. C.; and Collins, M. P., “One-
tudinally Reinforced Beams,” ACI Journal Proceedings, V. 81, No. 5, Way Shear Strength of Thick Slabs,” ACI Structural Journal, V. 103, No. 6,
Sept.-Oct. 1984, pp. 456-468. Nov.-Dec. 2006, pp. 794-802.
5. Vecchio, F. J., and Collins, M. P., “The Modified Compression-Field 19. Broms, B. B., “Crack Width and Crack Spacing in Reinforced
Theory for Reinforced Concrete Elements Subjected to Shear,” ACI Journal Concrete Members,” ACI Structural Journal, V. 62, No. 10, Oct. 1965,
Proceedings, V. 83, No. 2, Mar.-Apr. 1986, pp. 219-231. pp. 1237-1255.
6. Bažant, Z. P., and Yu, Q., “Design against Size Effect on Shear 20. Shioya, T., “Shear Properties of Large Reinforced Concrete
Strength of Reinforced Concrete Beams without Stirrups: II. Verification Member,” Special Report, Institute of Technology, Shimizu Corp., No. 25,
and Calibration,” Journal of Structural Engineering, ASCE, V. 131, No. 12, 1989, 198 pp.
2005, pp. 1886-1897. doi: 10.1061/(ASCE)0733-9445(2005)131:12(1886) 21. Shioya, T.; Iguro, M.; Nojiri, Y.; Akiyama, H.; and Okada, T., “Shear
7. Joint ACI-ASCE Committee 326, “Shear and Diagonal Tension,” ACI Strength of Large Reinforced Concrete Beams,” Fracture Mechanics:
Journal Proceedings, V. 59, No. 2, Feb. 1962, pp. 277-333. Application to Concrete, SP-118, V. C. Li and Z. P. Bažant, eds., American
8. ACI Committee 318, “Building Code Requirements for Reinforced Concrete Institute, Farmington Hills, MI, 1989, pp. 259-279.
Concrete (ACI 318-63),” American Concrete Institute, Farmington Hills, 22. Bentz, E. C., and Collins, M. P., “Development of the 2004 CSA
MI, 1963, 144 pp. A23.3 Shear Provisions for Reinforced Concrete,” Canadian Journal of
9. Murray, M. R., “An Investigation of the Unit Shear Strength of Civil Engineering, V. 33, No. 5, 2006, pp. 521-534. doi: 10.1139/l06-005
Geometrically Scaled Reinforced Concrete Beams,” thesis, Purdue Univer- 23. Sneed, L. H., 2007. “Influence of Member Depth on the Shear
sity, West Lafayette, IN, 2010, 106 pp. Strength of Concrete Beams,” PhD thesis, Purdue University, West Lafay-
10. Leonhardt, F., and Walther, R., “Contribution to the Treatment of ette, IN, 2007, 259 pp.
Shear in Reinforced Concrete,” Technical Translation 1172, J. P. Verschuren 24. Taylor, H. P. J., “Shear Strength of Large Beams,” Journal of the
and J. G. MacGregor, translators, 1962, 183 pp. Structural Division, ASCE, V. 98, Nov. 1972, pp. 2473-2490.
11. Iguro, M.; Shioya, T.; Nojiri, Y.; and Akiyama, H., “Experimental 25. American Concrete Institute, http://www.concrete.org. (last accessed
Studies on Shear Strength of Large Reinforced Concrete Beams under Jan. 30, 2015)
Uniformly Distributed Load,” Concrete Library of JSCE, translation from 26. McCain, K., “The Effect of Scale on the Resistance of Reinforced
Proceeding of JSCE, V. 1, No. 348, Aug. 1984, 18 pp. Concrete Beams to Shear,” thesis, Purdue University, West Lafayette, IN,
12. Frosch, R. J., “Behavior of Large-Scale Reinforced Concrete Beams 2012, 155 pp.
with Minimum Shear Reinforcement,” ACI Structural Journal, V. 97, 27. Daluga, D. R., “The Effect of Maximum Aggregate Size on the Shear
No. 6, Nov.-Dec. 2000, pp. 814-820. Strength of Geometrically Scaled Reinforced Concrete Beams,” thesis,
13. Sherwood, E. G.; Bentz, E. C.; and Collins, M. P., “Effect of Aggre- Purdue University, West Lafayette, IN, 2015, 200 pp.
gate Size on Beam-Shear Strength of Thick Slabs,” ACI Structural Journal, 28. ACI Committee 318, “Building Code Requirements for Structural
V. 104, No. 2, Mar.-Apr. 2007, pp. 180-190. Concrete (ACI 318-14) and Commentary (ACI 318R-14),” American
14. Sneed, L. H., and Ramirez, J. A., “Effect of Depth on the Shear Concrete Institute, Farmington Hills, MI, 2014, 519 pp.
Strength of Concrete Beams without Shear Reinforcement—Experimental 29. Bhal, N. S., “The Effect of Beam Depth on the Shear Capacity of
Study,” PCA R&D SN2921 Report, Portland Cement Association, Skokie, Single Span Reinforced Concrete Beams with and without Shear Reinforce-
IL, 2008, 182 pp. ment,” PhD thesis, University of Stuttgart, Stuttgart, Germany, 1968, 124 pp.
15. Yu, Q., and Bažant, Z. P., “Can Stirrups Suppress Size Effect on Shear
Strength of RC Beams?” Journal of Structural Engineering, ASCE, V. 137,
No. 5, 2011, pp. 607-617. doi: 10.1061/(ASCE)ST.1943-541X.0000295

14 ACI Structural Journal/January 2018


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 115-S01

Three-Dimensional Grid Strut-and-Tie Model Approach in


Structural Concrete Design
by Young Mook Yun, Byunghun Kim, and Julio A. Ramirez

The strut-and-tie model approach has now been incorporated and multiple strut-and-tie models may be required in design
in current U.S. design codes and guidelines for the design of situations involving multiple load combinations. Thus, even
disturbed regions in structural concrete elements. However, more in the case of 2-D design situations, the approach incorpo-
work is needed to extend the approach to design of three-dimen- rating stress trajectories presents significant challenges for
sional (3-D) structural concrete. It is also important to consider
efficient use in practice. Further, considering the case of
its verification with experimental evidence. The application to 3-D
a 3-D disturbed region with significant nonlinearity in the
design situations of this approach brings more uncertainties with
its proper application and limitations. To reduce uncertainty and distribution of stresses, the visualization of the stress trajec-
assist designers in the application of the strut-and-tie model to 3-D tories needed to deploy the strut-and-tie model is almost
situations, the authors present in this paper a 3-D grid strut-and-tie unattainable. Thus, in combination, the items described in
model approach consisting of three key steps: 1) grid elements to this paragraph could cause the improper placing of struts and
construct a 3-D strut-and-tie model; 2) triaxial failure model of ties and unsafe designs.
concrete to determine effective strengths of concrete struts and To overcome this problem, studies on the construction of
nodal zones; and 3) iterative technique to evaluate the axial stiff- 2-D and 3-D strut-and-tie models using performance-based
ness of struts and ties. In this paper, the authors also incorporate in and topology-based optimization techniques have been
the strut-and-tie model a new concept of maximum cross-sectional conducted (Liang et al. 2001; Ali and White 2001; Liang et
areas of struts and ties to examine the strut-and-tie model’s geomet-
al. 2002; Kwak and Noh 2006; Leu et al. 2006; Lee 2007).
rical compatibility. The approach is illustrated with the redesign
However, the construction of strut-and-tie models using
of a deep pile cap with tension piles available in the literature. In
a subsequent paper, the authors will evaluate the approach with optimization techniques has the following shortcomings: 1)
test results of 157 specimens tested to failure. The tests include 78 it can result in statically determinate strut-and-tie models
reinforced concrete pile caps, 19 slab-column joints, and 60 beams that induce difficulties in terms of distribution patterns of
subjected to torsion. reinforcing bars (which are different than practical hori-
zontal or vertical distributions of reinforcing bars, or both)
Keywords: disturbed region; effective strength; grid strut-and-tie model; and congestion of reinforcing bars due to large concen-
three-dimensional structural concrete.
trated tensile forces in the steel ties; 2) a fine finite element
(FE) mesh generation is required for initial FE modeling,
INTRODUCTION demanding significant time and effort; 3) different strut-
The strut-and-tie model approach for disturbed regions and-tie models are constructed according to the types of FE
(D-regions) in structural concrete promotes a better under- models and elimination criteria of FEs; and 4) FE models of
standing of load-transfer mechanisms and structural composite concrete members (concrete and reinforcing bars)
behavior, and it improves the designers’ ability to handle the are required to consider the proper structural behavior.
often complex and unusual design of these regions. These While use of statically determinate strut-and-tie models
advantages have resulted in the implementation of the strut- is advantageous to expedite structural analysis and subse-
and-tie model approach in major design codes and guidelines quent design, they may be limited in representing the load-
around the globe (CSA 2004; Comite Euro-International du transfer mechanisms of various types of 2-D and 3-D struc-
Beton 2010; American Association of State Highway and tural concrete disturbed regions. In addition, these models
Transportation Officials 2010; ACI Committee 318 2014). may limit the effective deployment of reinforcement due
Strut-and-tie model code approaches have been established to large tensile forces concentrated in the steel ties. These
by considering the behavior of two-dimensional (2-D) struc- limitations have led to studies on the use of statically inde-
tural concrete. Consequentially, their use in the analysis and terminate strut-and-tie models for the analysis and design of
design of three-dimensional (3-D) structural concrete with 2-D structural concrete (Alshegeir 1992; Foster and Gilbert
disturbed stress regions (for example, thick pile caps, bridge 1998; Hwang et al. 2000; Yun 2000; Reineck 2002; Tjhin
piers subject to multi-axial loads, and anchorage zones) is and Kuchma 2002; Bakir and Boduroglu 2005; MacGregor
often confusing.
The current practice of arranging the struts and ties based
ACI Structural Journal, V. 115, No. 1, January 2018.
on the load path method in accordance to the elastic stress MS No. S-2015-193.R2, doi: 10.14359/51700791, received February 28, 2017, and
trajectories can lead to several plausible models for the reviewed under Institute publication policies. Copyright © 2018, American Concrete
Institute. All rights reserved, including the making of copies unless permission is
same design situation. This situation can lead to potentially obtained from the copyright proprietors. Pertinent discussion including author’s
closure, if any, will be published ten months from this journal’s date if the discussion
unsafe outcomes, as it relies on the designer’s experience, is received within four months of the paper’s print publication.

ACI Structural Journal/January 2018 15


Fig. 1—Basic and auxiliary grid elements for construction of 3-D grid strut-and-tie model.
and Wight 2005; Park et al. 2005; Comite Euro-International situations, the authors present in this paper a 3-D grid strut-
du Beton 2010; Kim and Yun 2011). and-tie model approach.
Studies on the effective strength have been conducted based
on the theory of plasticity, FE nonlinear analyses of unrein- THREE-DIMENSIONAL GRID STRUT-AND-TIE
forced structural concrete, and failure mechanism of concrete MODEL
(Yun and Ramirez 1996; Yun 2000, 2005, 2006; Ha and Hong Construction of grid strut-and-tie model
2005). However, those studies were based on a proper place- A 3-D grid strut-and-tie model is composed of numerous
ment of concrete struts to describe the load transfer mecha- basic and auxiliary grid elements. It is constructed by
nism accurately and consider only 2-D situations. Thus, the considering the loading and geometrical conditions of the
uncertainty relating to the construction of a strut-and-tie structural concrete. A basic grid element with 12 horizontal/
model in 3-D situations has not been accounted for. vertical components, 12 inclined plane components, and
In this paper, a 3-D grid strut-and-tie model approach for four inclined space components is shown in Fig. 1(a). The
the analysis and design of structural concrete members with auxiliary grid elements, which are parts of the basic grid
D-regions is proposed. The proposed approach is deemed element, are shown in Fig. 1(b). The basic and auxiliary
sufficient to overcome the uncertainties and limitations grid elements allow for load transfers in all directions at a
of existing strut-and-tie model approaches related to the node. The auxiliary grid elements are used for structural
construction of the strut-and-tie model geometry, the selec- members that cannot be modeled properly by using basic
tion of the structural types of strut-and-tie model (determi- grid elements only, such as inclined pier caps, the top flanges
nate or indeterminate), and the determination of the effective of box girders subjected to lateral tendon forces, and flanges
strengths of struts and nodal zones. in T-shaped beams.
In the modeling of a load-transfer mechanism in structural
RESEARCH SIGNIFICANCE concrete with strut-and-tie models, the node of the model
The strut-and-tie model approach has now been incor- underneath a loading or bearing plate can be subjected to
porated in current U.S. design codes and guidelines for the a large concentrated force and possible crushing of the
design of disturbed regions in structural concrete elements. component (concrete strut) in a grid strut-and-tie model.
However, more work is needed to extend the approach to Thus, to prevent crushing of the concrete strut, the size of
the design of 3-D structural concrete. In this paper, a new a grid element must be selected by considering the sizes of
approach using the novel concept of grid strut-and-tie loading and bearing plates and the effective strength of the
models for application in 3-D design of structural concrete concrete strut. Further, the ratio of the vertical to horizontal
is presented. Its use is illustrated in the design of a pile cap, length of a grid element in two directions must be limited to
and the approach is verified with experimental evidence 0.75 ≤ lh/lvx ≤ 1.33 and 0.75 ≤ lh/lvy ≤ 1.33. In these relation-
in a subsequent paper. To reduce uncertainty and assist ships, lh, lvx, and lvx are the lengths of the boundary compo-
designers in the application of the strut-and-tie model to 3-D nents of a grid element (refer to Fig. 1). The background for

16 ACI Structural Journal/January 2018


Fig. 2—Procedure for determination of effective strengths of 3-D concrete struts.
this requirement is given in the “Geometrical compatibility bars. The procedure is illustrated using the example of a pile
of grid strut-and-tie model” section of this paper. The length cap shown in Fig. 2. The order of necessary steps is:
ratios limit the slopes of inclined struts between 28 and Step 1: From an FE model of 3-D unreinforced structural
62 degrees. concrete (Fig. 2(a)), select the FEs that are located closely
along the longitudinal axis of a concrete strut (Fig. 2(b)).
Effective strengths of struts and ties Step 2: Determine the principal stresses σk (k = 1 to 3,
A grid strut-and-tie model is composed of struts, ties, and σ1 ≤ σ2 ≤ σ3) and the principal directions (θ1, θ2) of each
nodal zones. In the strut-and-tie model design of structural FE after conducting a linear elastic FE analysis of the 3-D
concrete, the strength of struts and ties needs to be properly unreinforced structural concrete (Fig. 2(d)). By associating
determined. The design strength of a concrete strut, Pr,cs, the principal stresses σk with the failure criteria of 3-D unre-
may be expressed as inforced concrete described by the capped five-parameter
failure model (Willam and Warnke 1974), determine the
Pr,cs = ϕcs fcsAcs (1) failure principal stresses σkf (k = 1 to 3, σ1f ≤ σ2f ≤ σ3f) of
each finite element (Fig. 2(e)).
where ϕcs, fcs, and Acs are the strength reduction factor, the Step 3: Take σ1f as the effective strength fcs,e of a finite
effective strength, and the cross-sectional area of a concrete element if the principal direction of σ1f and the longitu-
strut, respectively. The effective strength of a concrete dinal axis of the concrete strut coincide. Reduce the effec-
strut is chosen as a fraction of the compressive strength of tive strength of the FE if its principal direction does not
concrete, fc′ coincide (Fig. 2(f)). This is accomplished by using a 3-D
Mohr’s circle (Fig. 2(g)) or by using the approach of axis
fcs = νfc′ (2) transformation. In the same manner, determine the effective
strengths of the other FEs located along the longitudinal axis
where ν is the coefficient of the effective strength of the of the concrete strut.
concrete strut. The coefficient is introduced to reflect the Step 4: Calculate the standard deviation of the effective
strength reduction of a concrete strut due to cracks parallel strengths of the FEs located closely along the concrete strut.
to the longitudinal axis of the strut. Then, determine the effective strength of concrete strut,
As previously discussed, current stress limits warrant fcs, by averaging the effective strengths of the FEs that are
further study for application to 3-D cases. In this paper, within the standard deviation range.
the method proposed by Yun (2005) for 2-D concrete struts Step 5: Evaluate the confinement force using the proce-
will be extended to 3-D applications. This method can be dure shown in Fig. 3 to reflect the effect of concrete confine-
used to determine the effective strengths of 3-D concrete ment by reinforcing bars. Apply the confinement force to the
struts consistently because it considers the state of stress, FE model of the 3-D unreinforced structural concrete along
the extent of cracking, the length of the struts, the deviation with external loads, and return to Step 2. Iterate Steps 2 to 5
angle between the strut axis and compressive principal stress until the difference of the confinement forces in previous and
flow, and the degree of confinement provided by reinforcing current steps is within a specified tolerance.

ACI Structural Journal/January 2018 17


respectively, are expressed as functions of their effective
strengths

Pr,st = ϕst fstAst; Pr,ct = ϕct fctAct (3)

where ϕst, fst, and Ast are the strength reduction factor of
the steel tie, the effective strength of the steel tie, and the
cross-sectional area of the steel tie, respectively. The vari-
able ϕst, fct (= 0.62√fc′), and Act are the strength reduction
factor of the concrete tie (which is the same as the concrete
strut’s strength reduction factor), the effective strength of
the concrete tie, and the cross-sectional area of the concrete
tie, respectively.

Effective strength of nodal zone


The cross-sectional forces of struts and ties must be trans-
ferred to neighboring struts and ties through nodal zones.
Accordingly, the load-carrying capacity of a strut-and-tie
model is controlled by the capacities of struts and ties and
by the capacities of nodal zones, requiring the verification
of the strengths of nodal zones. The strength of a nodal zone
Fig. 3—Algorithm for considering confinement effect by depends on factors such as: a) the confinement of the nodal
reinforcing bar. zone by reactions, compression struts, anchorage plates for
prestressing, reinforcement from the adjoining members,
Step 6: The effective strength of a concrete strut is influ-
and hoop reinforcing bars; b) the effects of strain discon-
enced by the compressive strength of the concrete (Berg-
tinuities within the nodal zone; and c) the splitting stresses
meister et al. 1993; MacGregor 1997). Thus, the effective
and hook-bearing stresses that result from the anchorage
strength fcs, determined in the previous step, is modified by
of the reinforcing bars in or immediately behind the nodal
multiplying the coefficients 1.1 – 0.25fc′/70, 1.0, and 0.85 for
zone. To consider the effects of the aforementioned factors
28 < fc′ < 70 MPa, fc′ ≤ 28 MPa, and fc′ ≤ 70 MPa,  respectively.
in the strength verification of nodal zones, the methods
In the strut-and-tie model analysis and design, the concrete
conducting FE material nonlinear analyses of unreinforced
struts are placed parallel to the directions of corresponding
concrete nodal zones (Yun 2006) or strut-and-tie models for
cracks or compressive principal stress trajectories. However,
nodal zones (Hong 2000) were introduced. Although the
it is difficult to satisfy this requirement. A generally accepted
two methods seem to be appropriate, they are not efficient
approach is to place a concrete strut as parallel as possible to
to implement in practice. Moreover, the methods are only
the elastically obtained compressive principal stress trajec-
applicable to the nodal zones in 2-D structural concrete.
tories. Here, angles of inclined struts and ties are controlled
In this paper, the approach in current design codes is used,
by the shapes of grid elements used in the modeling. This
verifying the strength of a nodal zone by comparing the
constraint may cause a difference in the directions between
maximum surface areas at the nodal zone boundaries with
the concrete struts and corresponding cracks, thus violating
the required surface areas. The maximum surface areas are
the requirement on the placement of concrete struts. In this
obtained by using the maximum cross-sectional areas of
study, the difference in the directions between the concrete
the struts and ties framing the nodal zone. The maximum
struts and corresponding cracks is compensated by adjusting
cross-sectional areas of struts and ties are defined in the
the effective strengths of concrete struts. If the direction of a
“Geometrical compatibility of grid strut-and-tie model”
concrete strut is parallel to the direction of a corresponding
section. The required surface areas at the nodal zone bound-
crack or the elastic compressive principal stress trajecto-
aries are obtained by dividing the cross-sectional forces of
ries, the concrete strut may carry the compressive force up
the struts and ties framing the nodal zone by the effective
to its effective strength. Otherwise, the effective strength of
strength of the nodal zone.
the concrete strut is reduced by considering the deviation
To examine the strengths of 3-D nodal zones, the effective
angles between the strut and the compressive principal stress
strengths of the nodal zones need to be established. Herein,
trajectories, as illustrated in Step 3 of Fig. 2. In addition, the
the effective strengths of the nodal zones were determined
value of 0.34fc′, suggested in ACI 318-14 for a concrete strut
by incorporating the capped five-parameter failure model
located at a tensile stress region, is taken as the minimum
of 3-D unreinforced concrete. The procedure is illustrated
effective strength in this study.
in Fig. 4. Because the strength of the 3-D nodal zone is
In the proposed approach, if tensile regions without rein-
generally greater than the strength of the 2-D nodal zone,
forcing bars exist in 3-D structural concrete, concrete ties
the effective strength of the ACI 318-14 Compression-Com-
are used to describe an appropriate load-transfer mechanism.
pression-Tension (C-C-T) nodal zone, 0.51fc′, was taken as
The design strengths of steel and concrete ties, Pr,st and Pr,ct,
the minimum strength in this study.

18 ACI Structural Journal/January 2018


Fig. 4—Procedure for determination of effective strength of 3-D nodal zones. (Note: 1 MPa = 145 psi.)

Fig. 5—Algorithm for determining axial rigidities of struts and ties.


Stiffness of struts and ties model. The algorithm for determining the cross-sectional
The required cross-sectional areas of struts and ties in a areas of struts and ties is given in Fig. 5.
grid strut-and-tie model under design loads were determined Once the stiffness of concrete struts and steel ties is
using an iterative technique. The approach identifies the obtained multiplying the cross-sectional areas of the struts
main struts and ties that play an important role in transfer- and ties by their modulus of elasticity, the indeterminate
ring design loads to supports or neighboring members, and grid strut-and-tie model must be analyzed to determine the
assigns appropriate stiffness of struts and ties to enable the concrete struts and steel ties. In this analysis, an inclined
structural analysis of an indeterminate grid strut-and-tie plane component crossing an inclined plane concrete strut

ACI Structural Journal/January 2018 19


Fig. 6—Maximum cross-sectional areas of components of basic grid element.
placed almost parallel to the compressive stress trajectory maximum areas that struts and ties can have without over-
by transferring a large force would generally be consid- laps in a grid element (refer to Fig. 6).
ered a steel tie. In this case, the cross-sectional force of the The maximum cross-sectional area of a component of a
steel tie can be replaced with 2-D horizontal and vertical grid element is determined by multiplying two maximum
forces. The same rule applies to cases of irregularly-shaped widths of the component evaluated on the component’s two
grid elements. orthogonal planes. For example, as shown in Fig. 6(c), the
maximum cross-sectional area of the x-direction component
Geometrical compatibility of grid strut-and-tie of length lx1 is determined by multiplying the z-direction
model maximum width wHE and the y-direction maximum width
The geometrical compatibility of a 3-D grid strut-and-tie wVE. Here, the maximum width wHE is taken as 0.2lz1 + 0.2lz2,
model must be examined as part of the strut-and-tie model with lz1 and lz2, corresponding respectively to the lengths
design. The required cross-sectional areas of struts and wH1 and wH1 shown in Fig. 6(a). These are the lengths of
ties, determined using an iterative approach, must be less the z-direction components connected to the ends of the
than their maximum cross-sectional areas. The maximum x-direction component. The reason for taking 20% of the
cross-sectional areas of struts and steel ties in the strut- lengths of the neighboring elements is based on the rule
and-tie model application in current design codes are deter- that the maximum widths of all elements in a grid element
mined by using initial design conditions such as the geom- should be the same initially. A mathematical derivation for
etry of the structural concrete, the sizes of bearing and the 20% rule is provided in the Appendix of this paper. Simi-
loading plates, and the pattern of reinforcement details. In larly, the maximum width wVE is defined as 0.2ly1 + 0.2ly2,
the proposed approach, the method employed in the 2-D where ly1 and ly2 are the lengths of the y-direction compo-
grid strut-and-tie model approach proposed by Yun and Kim nents connected to the ends of the x-direction component.
(2008) is extended. Namely, in addition to the strut-and-tie The maximum cross-sectional areas of the y- and z-direction
model approaches of current design codes, the maximum components are determined following a similar approach.
cross-sectional areas of struts and ties are defined by the The maximum cross-sectional area of an inclined plane
component of a given grid element is determined in the same

20 ACI Structural Journal/January 2018


manner. As shown in Fig. 6(d), the maximum cross-sectional
area of the inclined plane component of x-direction length lx1
in the xy-plane is determined by multiplying the maximum
width wIPE of the xy-plane and the maximum width wHEZ of
the plane orthogonal to the xy-plane. Here, the maximum
width wIPE of a given plane is determined following the
method illustrated in Fig. 6(b). The maximum width wIPE
of the xz-plane is taken as the length (the smaller of wa +
wb and wc + wd) of the perpendicular lines connecting the
points (a, b, c, d), which are determined by considering the
maximum widths of the neighboring horizontal and vertical
elements. Following the methods illustrated in Fig. 6(a) and
6(b), the maximum widths wIPE and wHEZ of the xy-plane and
its orthogonal planes, respectively, are expressed as

0.6lx1l y 2 Fig. 7—Modified maximum cross-sectional areas of compo-


wIPE = , wHEZ = 0.2ls1 + 0.2ls 2 (4) nents of basic grid element.
lx21 + l y22
nent CA exceeds the maximum cross-sectional area defined
where lx1, ly2, lz1, and lz2, as shown in Fig. 6(d), are the lengths above, the maximum cross-sectional area of the component
of the grid element components. The maximum cross-sec- AIPE (= wIPE × wHEZ) increases to AIPE (= wIPE × wHEZ ) if the
tional areas of the inclined plane components in the xz-plane required widths of the neighboring components CF~CI of the
and yz-plane are determined following a similar approach. xy-plane and CB~CE of the plane orthogonal to the xy-plane
The cross-sectional area of an inclined space component are less than their maximum widths. In Fig. 7, w′HEZ is the
that exists only in a 3-D grid element is defined as the area limit of the modified maximum width of component CA that
excluding all the cross-sectional areas of horizontal compo- can be maximized when the required widths of the compo-
nents, vertical components, and inclined plane components nents CB~CE are equal to zero. Also, w′IPE is the limit of
in a grid element. For example, as shown in Fig. 6(e), the the modified maximum width of component CA that can be
maximum cross-sectional area of the space component maximized in the xy-plane. The concept employed herein
is determined by multiplying the maximum width wIPE of originates from the rule stating that the overlaps of areas
the xz-plane and the maximum width wISE of the yz-plane (or widths) of neighboring components are not allowed,
rotated about the y-axis. The maximum widths wIPE and wISE as introduced in Fig. 6. The modified maximum width of a
are expressed as component of a grid element in any plane of a 3-D grid strut-
and-tie model is obtained in a similar manner.
0.6lx1lz 2 0.6 lx21 + lz22 l y 2
wIPE = , wISE = (5) IMPLEMENTATION
lx21 + lz22 lx21 + l y22 + lz22 The analysis or design of D-Regions in structural concrete
using the proposed approach incorporating basic grid
The maximum cross-sectional areas of the other three elements to construct a 3-D strut-and-tie model is conducted
space components are determined in the same way. Fig. 6(f) using the procedure shown in Fig. 8. A detailed explanation
shows an example where the entire available volume of of the procedure follows.
concrete is required when the maximum cross-sectional areas Step 1: Set up the initial conditions, including loading
of all components in a 3-D grid strut-and-tie model are used and boundary conditions, material strengths, and sizes of
without overlaps. the loading and bearing plates. For a strut-and-tie model anal-
In the proposed approach, a component of a grid element ysis using the proposed approach, information regarding
is assumed to fail if its maximum cross-sectional area is less reinforcement details is required to calculate the cross-
than that required for carrying the cross-sectional force. To sectional areas of the steel ties.
reflect more precisely the characteristics of the load-transfer Step 2: Conduct a linear elastic FE analysis. The concrete
mechanism in a grid element, the modified maximum element or structure is modeled as a plain (unreinforced)
cross-sectional area of a component that is determined by isotropic 3-D problem. Construct a 3-D grid strut-and-tie
considering the required cross-sectional areas of neigh- model using basic and auxiliary grid elements. To facilitate
boring components (instead of the maximum cross-sectional a proper load transfer within the concrete structure, the size
area defined above) is used for examining the geometrical of grid elements must be selected considering the sugges-
compatibility. Here, the modified maximum cross-sectional tions given in the section, “Construction of grid strut-and-tie
area of a component is determined by multiplying two modi- model.”
fied maximum widths of the component in two orthogonal Step 3: Determine the effective strengths of the concrete
planes. An example of determining the modified maximum struts by using the linear elastic FE analysis results.
area of an inclined plane component of a grid element is Step 4: Determine the axial stiffness and cross-sectional
shown in Fig. 7. If the required cross-sectional area of compo- forces of the struts and ties using the iterative technique

ACI Structural Journal/January 2018 21


Fig. 8—Procedure for applying 3-D grid strut-and-tie approach.
shown in Fig. 5. For strut-and-tie model analysis, the stiff- Step 7: In this linear analysis, the ultimate strength of the
ness of the steel ties must be determined using the informa- structural concrete is estimated by examining the load-car-
tion on reinforcement details given in Step 1. Determine the rying capacities of the struts, ties, and nodal zones of the
effective strengths of the 3-D nodal zones using the linear constructed 3-D grid strut-and-tie model under its ultimate
elastic FE analysis results of the grid strut-and-tie model. loads while satisfying strut-and-tie model’s geometrical
Step 5: To consider the effects of the confinement of compatibility condition. The ultimate strength is obtained by
concrete by reinforcing bars on the effective strengths of examining the conditions given in Table 1 for compressive
concrete struts, apply the cross-sectional forces of the hori- failure of the concrete struts, tensile failure of horizontal (or
zontal and vertical steel ties additionally to the FE model of vertical) concrete ties, tensile failure of steel ties, and local
the structural concrete. Iterate Steps 2 to 5 until the cross-sec- failure of nodal zones. Here, failures of the inclined plane
tional forces of the horizontal and vertical steel ties converge. and space concrete ties of a grid element were excluded
Step 6: Determine the maximum cross-sectional areas of because the cross-sectional forces are distributed to the
the struts and ties to verify the conditions of strut-and-tie nearby horizontal and vertical ties. In the nonlinear strut-
model’s geometrical compatibility and nodal zone strength. and-tie model analysis, the ultimate strength of the struc-
If necessary, the maximum cross-sectional areas of the struts tural concrete is estimated by examining the condition of
and ties should be adjusted. structural stability of the constructed 3-D grid strut-and-tie

22 ACI Structural Journal/January 2018


Table 1—Ultimate strength defined by conditions of 3-D grid strut-and-tie model’s components
Component Ultimate strength (= Pu) States of components at Pmax
Concrete strut Pmax Acs req’d = Acs pro’d, Act req’d < Act pro’d, fs < fy, σnz < fnz
Concrete tie Pmax Acs req’d < Acs pro’d, Act req’d = Act pro’d, fs < fy, σnz < fnz
Steel tie Pmax × fy ÷ fs Acs req’d ≤ Acs pro’d, Act req’d ≤ Act pro’d, fs > fy, σnz < fnz
Nodal zone Pmax × fnz ÷ σnz Acs req’d ≤ Acs pro’d, Act req’d ≤ Act pro’d, fs < fy, σnz > fnz

Notes: Pmax is maximum load that 3-D grid strut-and-tie model can carry by satisfying condition of strut-and-tie model’s geometrical compatibility (Acs req’d ≤ Acs pro’d and Act
req’d ≤ Act pro’d); Acs req’d and Act req’d are required cross-sectional areas of concrete strut and concrete tie at Pmax, respectively; Acs pro’d and Act pro’d are maximum cross-sectional areas of
concrete strut and concrete tie, respectively; fy and fs are yield strength of steel and stress of steel tie at Pmax, respectively; fnz is effective strength of nodal zone; σnz is compressive
stress of nodal zone boundary at Pmax.

Fig. 9—Geometry and load condition of pile cap. (Note: 1 mm = 0.0394 in.; 1 MPa = 0.145 ksi; 1 kN = 0.225 kip.)
model, as shown in Fig. 8. The conditions of the strut-and-tie struts and nodes, and anchorage and reinforcing bar details
model’s geometrical compatibility and nodal zone strength are described in the reference.
must be satisfied. If this is not the case, it will be necessary To design the pile cap by the proposed approach, a linear
to iterate, by changing the initial conditions, or increasing elastic 3-D solid finite element analysis about the pile cap
the strength of the selected model, to satisfy the original is conducted first. In the analysis, pile reactions are applied
conditions. Once the initial conditions are satisfied, the rein- as external forces as in the case of ACI SP-273. A hinge
forcement is selected to satisfy the calculated tie forces at the and multiple rollers at the interface of column and pile cap
proper locations. are imposed to describe the structural behavior of the pile
cap properly and stabilize the finite element model. Next,
DESIGN OF DEEP PILE CAP WITH TENSION PILES a 3-D grid strut-and-tie model is constructed. To compare
The deep pile cap with tension piles, used in the illus- the design results, the top and bottom nodes of the grid
tration of the ACI 318 strut-and-tie model method in ACI strut-and-tie model are located at 75 and 200 mm (2.95 and
SP-273 (Reineck and Novak 2010), is selected to illustrate 7.87 in.) away from the top and bottom surfaces of the pile
the grid strut-and-tie model design procedure and compare cap, as in the case of the ACI SP-273 strut-and-tie model.
the results with those obtained in the original example. The The x- and z-coordinates of the nodes underneath the column
geometry of pile cap, material properties, loading, and geom- and at pile supports in the grid strut-and-tie model are the
etry of a 3-D strut-and-tie model constructed in ACI SP-273 same as those of the ACI SP-273 strut-and-tie model. The
are shown in Fig. 9. Details regarding the load condition, number of grids is decided to satisfy the strength require-
construction of strut-and-tie model geometry by conven- ment at critical nodes (such as the nodes at supports and
tional strut-and-tie model approach, strength verifications of underneath column) and describe the principal compressive
stress flows. The shapes of grid elements are constructed to

ACI Structural Journal/January 2018 23


fulfill the suggested requirement on the vertical to horizontal
length ratios of a grid element in x- and z-directions.
The effective strengths of concrete struts and nodal zones
are determined by following the procedures explained in
Fig. 2, 3, and 4. The cross-sectional areas of struts and ties
are determined by using the iterative technique shown in
Fig. 5. Dimensioned shapes of the grid strut-and-tie model
are shown in Fig. 10. Based on the dimensioned shapes satis-
fying the geometrical compatibility of strut-and-tie model, it
is known that the strength conditions of concrete struts and
nodes are met.
The cross-sectional forces of ties in the ACI SP-273 and
grid strut-and-tie models are determined and shown in Fig.
11. Using the tie forces, the required areas of steel are deter-
mined. The required areas of reinforcement by the two types
of strut-and-tie models are in reasonably good agreement
at every region of the pile cap. However, the total required
areas of reinforcing bars by the proposed approach are 9%
less compared with that of the ACI SP-273 approach, indi-
cating that the strut-and-tie model constructed using the
proposed approach is more efficient because it undergoes
Fig. 10—Dimensioned shapes of 3-D grid strut-and-tie less strain energy. This is achieved by placing the struts and
model for pile cap. (Note: Full-color rendering of figure can ties according to principal stress trajectories.
be viewed online at www.concrete.org.)

Fig. 11—Cross-sectional forces of ties in pile cap strut-and-tie models. (Note: 1 kN = 0.225 kip.)

24 ACI Structural Journal/January 2018


SUMMARY Bakir, P. G., and Boduroglu, H. M., 2005, “Mechanical Behavior and
Non-linear Analysis of Short Beams using Softened Truss and Direct Strut
The current strut-and-tie model approach in U.S. guide- & Tie Models,” Engineering Structures, V. 27, No. 4, pp. 639-651. doi:
lines is implemented using statically determinate models. 10.1016/j.engstruct.2004.12.003
While in many cases, determinate strut-and-tie models are Bergmeister, K.; Breen, J. E.; Jirsa, J. O.; and Kreger, M. E., 1993,
“Detailing in Structural Concrete,” Research Report 1127-3F, University of
sufficient, they have limits in representing the load-transfer Texas at Austin, Austin, TX.
mechanisms in various types of structural concrete with Canadian Standards Association, 2004, “Design of Concrete Structures
disturbed regions. The work presented herein is a refinement for Buildings (A23.3-M04),” Rexdale, ON, Canada.
Comite Euro-International du Beton, 2010, “CEP-FIP Model Code
of the strut-and-tie model approach in the construction of 2010,” International Federation for Structural Concrete (fib), Lausanne,
the model itself, the selection of the model’s structural type, Switzerland.
and in the determination of the effective strengths of struts Foster, S. J., and Gilbert, R. I., 1998, “Experimental Studies on High-
Strength Concrete Deep Beams,” ACI Structural Journal, V. 95, No. 4,
and nodal zones. The proposed 3-D grid strut-and-tie model July-Aug., pp. 382-390.
approach is intended to extend the application of the strut- Ha, T. H., and Hong, S. G., 2005, “Effect of Diagonal Cracking on
and-tie approach by illustrating the implementation of 3-D the Strength of Concrete Strut in RC Members,” Journal of the Korean
Concrete Institute, V. 17, No. 1, pp. 383-386.
strut-and-tie models to complex design situation that are more Hong, S. G., 2000, “Strut-and-Tie Models and Failure Mechanisms for
realistically represented by 3-D load-transfer mechanisms in Bar Development in Tension-Tension-Compression Nodal Zone,” ACI
structural concrete. It also illustrates the implementation of Structural Journal, V. 97, No. 1, Jan.-Feb., pp. 111-121.
Hwang, S. J.; Lu, W. Y.; and Lee, H. J., 2000, “Shear Strength Predic-
a computer-based method aimed at eliminating numerous tion for Deep Beams,” ACI Structural Journal, V. 97, No. 3, May-June,
hand calculations and simplifying the strength verification pp. 367-376.
in the strength analysis and design of the entire 3-D member. Kim, B. H., and Yun, Y. M., 2011, “An Indeterminate Strut-and-tie
Model and Load Distribution Ratio for RC Deep Beams - (I) Model & Load
The implementation of the proposed approach requires first Distribution Ratio,” Advances in Structural Engineering, V. 14, No. 6,
a FE linear analysis of the member using a strut-and-tie pp. 1031-1041. doi: 10.1260/1369-4332.14.6.1031
model of the member without reinforcement. The proposed Kwak, H. G., and Noh, S. H., 2006, “Determination of Strut-and-Tie
Models Using Evolutionary Structural Optimization,” Engineering Struc-
approach using a 3-D grid strut-and-tie model is presented as tures, V. 28, No. 10, pp. 1440-1449. doi: 10.1016/j.engstruct.2006.01.013
an effective and practical design technique offering a more Lee, W. S., 2007, “Strut-and-Tie Model Approach Using Evolutionary
effective alternative to the nonlinear finite element analysis Structural Optimization for Reinforced Concrete Structural Members,”
PhD dissertation, Kyungpook National University, Daegu, Korea, 210 pp.
for the design, and particularly the detailing, of structural Leu, L. J.; Huang, H. W.; Chen, C. S.; and Liao, Y. P., 2006, “Strut-
concrete members and subassemblies. and-Tie Design Methodology for Three-Dimensional Reinforced Concrete
In a subsequent paper, the ultimate strength of 78 rein- Structures,” Journal of Structural Engineering, ASCE, V. 132, No. 6, pp.
929-938. doi: 10.1061/(ASCE)0733-9445(2006)132:6(929)
forced concrete pile caps, 19 slab-column joints, and 60 Liang, Q. Q.; Uy, B.; and Steven, G. P., 2002, “Performance-based Opti-
torsional beams have been evaluated to verify the proposed mization for Strut-and-Tie Modeling of Structural Concrete,” Journal of
approach in this paper. In addition, the design of a pier cap Structural Engineering, ASCE, V. 128, No. 6, pp. 815-823. doi: 10.1061/
(ASCE)0733-9445(2002)128:6(815)
subjected to multiple load combinations with longitudinal Liang, Q. Q.; Xie, Y. M.; and Steven, G. P., 2001, “Generating Optimal
and lateral loads is also conducted to illustrate its application. Strut-and-Tie Models in Prestressed Concrete Beams by Performance-based
Optimization,” ACI Structural Journal, V. 98, No. 2, Mar.-Apr., pp. 226-232.
MacGregor, J. G., 1997, Reinforced Concrete: Mechanics and Design,
AUTHOR BIOS third edition, Prentice Hall, Englewood Cliffs, NJ.
Young Mook Yun is a Professor in the Department of Civil Engineering
MacGregor, J. G., and Wight, J. K., 2005, Reinforced Concrete:
at Kyungpook National University, Daegu, Korea. He received his PhD
Mechanics and Design, fourth edition, Pearson Prentice Hall, Upper Saddle
from Purdue University, West Lafayette, IN, in 1995. His research inter-
River, NJ.
ests include strut-and-tie model analysis and design of concrete members,
Park, H. G.; Kim, Y. G.; and Eom, T. S., 2005, “Direct Inelastic Strut-
computational mechanics, and computer graphics.
and-tie Model Using Secant Stiffness,” Journal of the Korean Concrete
Institute, V. 17, No. 2, pp. 201-212. doi: 10.4334/JKCI.2005.17.2.201
Byunghun Kim is a Lead Engineer in the Infrastructure Department
Reineck, K. H., ed., 2002, Examples for the Design for Structural
of Hyundai Engineering, Seoul, South Korea. He received his PhD from
Concrete with Strut-and-Tie Models, SP-208, American Concrete Institute,
Kyungpook National University in 2004. His research interests include
Farmington Hills, MI.
strut-and-tie model design of concrete structures and development of
Reineck, K. H., and Novak, L. C., eds., 2010, Further Examples for the
numerical analysis methods for steel and reinforced concrete structures.
Design of Structural Concrete with Strut-and-Tie Models, SP-273, Amer-
ican Concrete Institute, Farmington Hills, MI, 288 pp.
Julio A. Ramirez, FACI, is a Professor in the Lyles School of Civil Engi-
Tjhin, T. N., and Kuchma, D. A., 2002, “Computer-Based Tools for
neering at Purdue University. He is a member of Joint ACI-ASCE Commit-
Design by Strut-and-Tie Method: Advances and Challenges,” ACI Struc-
tees 408, Bond and Development of Reinforcement; and 445, Shear and
tural Journal, V. 99, No. 5, Sept.-Oct., pp. 586-594.
Torsion. He received the ACI Delmar L. Bloem Distinguished Service
Willam, K. J., and Warnke, E. P., 1974, “Constitutive Model for the
Award and the ACI Joe W. Kelly Award in 2000 and 2006, respectively.
Triaxial Behavior of Concrete,” Proceedings of the International Associa-
tion of Bridge Structures, V. 19, pp. 1-30.
REFERENCES Yun, Y. M., 2000, “Nonlinear Strut-and-tie Model Approach for Structural
ACI Committee 318, 2014, “Building Code Requirements for Struc- Concrete,” ACI Structural Journal, V. 97, No. 4, July-Aug., pp. 581-590.
tural Concrete (ACI 318-14) and Commentary (ACI 318R-14),” American Yun, Y. M., 2005, “Effective Strength of Concrete Strut in Strut-and-tie
Concrete Institute, Farmington Hills, MI, 520 pp. Model (I): Methods for Determining Effective Strength of Concrete Strut,”
Ali, M. A., and White, R. N., 2001, “Automatic Generation of Truss Journal of the Korean Society of Civil Engineers, V. 25, pp. 49-59.
Model for Optimal Design of Reinforced Concrete Structures,” ACI Struc- Yun, Y. M., 2006, “Strength of Two-Dimensional Nodal Zones in Strut-
tural Journal, V. 98, No. 4, July-Aug., pp. 431-442. and-tie Models,” Journal of Structural Engineering, ASCE, V. 132, No. 11,
Alshegeir, A., 1992, “Analysis and Design of Disturbed Regions with pp. 1764-1783. doi: 10.1061/(ASCE)0733-9445(2006)132:11(1764)
Strut-and-tie Models,” PhD dissertation, School of Civil Engineering, Yun, Y. M., and Kim, B. H., 2008, “Two-Dimensional Grid
Purdue University, West Lafayette, IN. Strut-Tie Model Approach for Structural Concrete,” Journal of Struc-
American Association of State Highway and Transportation Officials, tural Engineering, ASCE, V. 134, No. 7, pp. 1199-1214. doi: 10.1061/
2010, “AASHTO LRFD Bridge Design Specifications,” fifth edition, Wash- (ASCE)0733-9445(2008)134:7(1199)
ington, DC.

ACI Structural Journal/January 2018 25


Yun, Y. M., and Ramirez, J. A.1996 , “Strength of Struts and Nodes in
Strut-and-Tie Model,” Journal of Structural Engineering, ASCE, V. 122,
No. 1, pp. 20-29. doi: 10.1061/(ASCE)0733-9445(1996)122:1(20)

APPENDIX
Mathematical proof for taking 20% of the length of
neighboring horizontal (vertical) components in the
determination of the maximum width of vertical (horizontal)
components (Fig. A1).
By using the equation of perpendicular, the half of the
maximum width of the inclined plane component of the
square grid element shown previously is described as follows

ax3 + by3 + c
wb = (A1)
(a 2 + b 2 )
Fig. A1.
where the coefficients a, b, and c are a = y1 – y2 = –1.0, b = x2
– x1 = 1.0, and c = x1y2 – y1x2 = 0, respectively. Substituting maximum widths of the vertical and inclined components
the coefficients into Eq. (A1), we have are the same, that is, by equating wa to wb, we have

− wa + 0.5 (A2) 0.5


wb = wa = = 0.2071 ≈ 0.2 (A3)
2 2 +1

As the value 0.5 is the maximum half width that the
vertical component can take, that is, as wa = 0.5, |–wa + Therefore, the maximum width of the inclined plane
0.5| becomes –wa + 0.5. By assuming that the halves of the component equals to those of neighboring vertical
(horizontal) components if we take 20% of the length of
neighboring horizontal (vertical) components.

26 ACI Structural Journal/January 2018


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 115-S02

Verification of Three-Dimensional Grid Strut-and-Tie Model


Approach in Structural Concrete
by Young Mook Yun, Hyun Soo Chae, Byunghun Kim, and Julio A. Ramirez
In this study, a proposed three-dimensional (3-D) grid strut-and-tie also examined through the design of a pier cap subjected
model approach is verified using the experimental capacity from to multiple load combinations with longitudinal and lateral
78 reinforced concrete pile caps, 19 slab-column joints, and 60 loads. The design results of the pier cap were compared with
torsional beams obtained by others. The analysis results were the results using the ACI 318 sectional design methods. The
compared with those obtained using the BS 8110, EC2, CRSI,
reliability of the design results were verified using the strut-
fib, AASHTO-LRFD, and ACI 318 design provisions. In addition,
and-tie models suggested in current design codes.
the design of a pier cap subjected to multiple load combinations
with longitudinal and lateral loads was conducted to illustrate the
proposed approach. The design results from the pier cap example RESEARCH SIGNIFICANCE
were also compared with those obtained using ACI 318-14. The To overcome limitations and improve clarity in the design
analysis results agreed well with experimental results. of structural concrete using 3-D strut-and-tie model approach
in current design codes, the authors have proposed a 3-D
Keywords: grid strut-and-tie model; pier cap; pile cap; slab-column joint; grid strut-and-tie model approach (Yun et al. 2017). The
torsional beam.
approach consists of three key elements: 1) use of basic grid
elements to construct a 3-D strut-and-tie model; 2) a triaxial
INTRODUCTION failure model of concrete to determine effective strengths of
In the paper by Yun et al. (2017), the limitations of the concrete struts and nodal zones; and 3) a simple iterative
strut-and-tie model approach in current U.S. design codes technique to evaluate the axial stiffness of struts and ties. In
and guidelines with respect to the construction of the strut- this paper, the approach is validated using an extensive data-
and-tie model, selection of the strut-and-tie model’s struc- base of relevant tests including reinforced concrete pile caps,
tural type, and determination of the effective strengths of slab-column joints, and torsional beams. In this verification
concrete struts and nodal zones for the designs of three- of the approach the provisions of BS 8110, EC 2, CRSI, fib,
dimensional (3-D) structural concrete with D-regions were AASHTO-LRFD, and ACI 318-14 were used. Finally, the
examined. To overcome the limitations of the strut-and-tie approach is illustrated using an example of a bridge pier cap
model approach in U.S. codes of practice, a 3-D grid strut- under multiple loading conditions.
and-tie model approach that incorporates grid elements
to construct a 3-D strut-and-tie model was proposed. The ANALYSIS OF PILE CAP TESTS
approach described in the paper by Yun et al. (2017) includes To verify the 3-D grid strut-and-tie model approach, the
a five-parameter failure model of unreinforced concrete to ultimate strengths of 78 reinforced concrete pile caps, tested
determine the effective strengths of concrete struts and nodal to failure by Clarke (1973), Sabnis and Gogate (1984),
zones and an iterative technique to evaluate the axial stiff- and Suzuki et al. (1998, 2000) were estimated using the
ness of struts and ties. The maximum cross-sectional area of approach. In addition, capacities were calculated using ACI
struts and ties approach is used to examine the strut-and-tie 318 sectional design methods (1999, 2014), CRSI sectional
model’s geometrical compatibility with the strut-and-tie design method (2008), and ACI 318-14 strut-and-tie model
model approaches of current design codes. approach (2014). The pile cap tests were conducted to inves-
The proposed approach was validated using the experi- tigate the effects of patterns of flexural reinforcing bars on
mental capacity of 78 reinforced concrete pile caps, 19 slab- their structural behaviors and strengths (Clarke 1973; Sabnis
column joints, and 60 torsional beams obtained by others. and Gogate 1984; and Suzuki et al. 1998, 2000). The mate-
The ultimate strength of 78 pile caps estimated using the rial properties and geometries of the pile caps are summa-
proposed approach was compared with ultimate strength rized in Table 1. Reinforcement details and geometric shapes
based on provisions in ACI 318 sectional design methods of representative pile caps tested by Suzuki et al. (1998)
(1999, 2014), CRSI sectional design method (2008), and are shown in Fig. 1. The pile caps were failed by bending
ACI 318-14 strut-and-tie model approach (2014). The ulti- (flexure) and/or shear (one-way, two-way, and punching).
mate strengths of the slab-column joints predicted using Detailed information, including the failure patterns of the
the proposed approach were compared with those from the pile caps, is given in the aforementioned references.
two-way slab provisions of BS 8110 (1997), EC2 (2004), fib
(2010), and ACI 318-14 (2014). The ultimate strengths of ACI Structural Journal, V. 115, No. 1, January 2018.
MS No. S-2015-192.R2, doi: 10.14359/51700946, received July 17, 2015, and
the torsional beams were compared with those of AASHTO- reviewed under Institute publication policies. Copyright © 2018, American Concrete
LRFD (2010), ACI 318-14 (2014), Hsu and Mo (1983), Institute. All rights reserved, including the making of copies unless permission is
obtained from the copyright proprietors. Pertinent discussion including author’s
and Rahal and Collins (1995). The proposed approach was closure, if any, will be published ten months from this journal’s date if the discussion
is received within four months of the paper’s print publication.

ACI Structural Journal/January 2018 27


Table 1—Specification of pile caps
Reinforcing
Investigators (A) w/d d, mm Column size, mm Pile size, mm fc′, MPa fy, MPa ρ, % bar layout
Clarke (1973) 12 0.25 to 0.49 109 to 117 200, S 200, R 17.9 to 41.0 510 0.188 to 0.238 B and G
Sabnis and Gogate (1984) 8 0.54 to 0.58 400 76, R 76, R 17.4 to 36.3 414 0.136 to 1.287 G
Suzuki et al. (1998) 28 0.40 to 0.80 150 to 250 300, S 150, R 18.9 to 31.3 356 to 413 0.228 to 0.546 B and G
Suzuki et al. (2000) 30 0.29 to 0.67 150 to 350 200 to 300, S 150, R 24.6 to 28.8 358 to 383 0.181 to 0.272 G
Total 78 0.25 to 0.80 109 to 400 76 to 300 76 to 200, R 17.4 to 41.0 356 to 510 0.136 to 1.287 B, G

Notes: (A) is number of pile caps; w is distance from face of column to center of nearest pile; d is effective depth; ρ is flexural reinforcement ratio; R, S, B, and G is round, square,
bunched, and grid; 1 mm = 0.0394 in.; 1 MPa = 145 psi.

Fig. 1—Reinforcement details and geometrical shapes of pile caps (Suzuki et al. 1998). (Note: 1 mm = 0.0394 in.)
The ultimate strength of the pile caps based on ACI 318 the nominal shear strength was determined from ACI 318-99
sectional design methods was determined by applying the Eq. (11.29).
flexural design provisions of beams and slabs (ACI 318-14, The detailed procedure for evaluating the pile cap’s
Sections 22.2 through 22.4), the bearing strength provisions ultimate strength is shown in Tables 2(a) and (b) with an
(ACI 318-14, Section 10.14), the shear design provisions of example of the specimen BPC-30-25-1 (Suzuki et al. 1998).
one-way and two-way slabs (ACI 318-14, Sections 22.5 and In an application of the ACI 318 sectional design method,
22.6), and the shear design provisions of deep beams with the minimum value among Pu, f, Pu,bc, Pu,bp, Pu,1s, and Pu,2s
ln/d ≤ 5 (where ln is the length of clear span, and d is the evaluated, respectively, from the flexural strength Mn, f, the
distance from the extreme compression fiber to the center bearing strength Pn,bc at column, the bearing strength Pn,bp
of longitudinal reinforcement) (ACI 318-99, Section 11.8). at the pile, the shear strength of one-way slab Vn,1s, and the
In the one-way shear design of the pile cap, the nominal shear strength of two-way slab shear Vn,2s, was taken as the
shear strength provided by the concrete at the critical section ultimate strength of the pile cap Pu. The ultimate strengths of
was determined from ACI 318-14 Eq. (22.5.5.1) and Eq.(a) the other pile caps were determined in the same way.
of ACI 318-14 Table 22.5.5.1. For the two-way shear design, The CRSI sectional design method for the pile caps is
the section along the circumference separated from the face quite similar to the ACI 318 sectional design method. For the
of the column by a distance of d/2 was regarded as the crit- shear design of a pile cap that can be regarded as a one-way
ical section, and the smallest value determined from Eq.(a) slab, the section at the face of the column is considered to be
to (c) of ACI 318-14 Table 22.6.5.2 was taken as the nominal the critical section, and the nominal shear strength Vc at the
shear strength at the critical section. When the ratio of the critical section is determined from the following equations
clear span to the effective depth of pile cap was less than 5,
Vc = VcACI ≥ 0.17 f c′bd for w/d ≥ 1.0 (1a)

28 ACI Structural Journal/January 2018


Table 2(a)—Evaluation of ultimate strength of pile cap BPC-30-25-1 by current design codes:
sectional methods
Bearing
Methods Flexure Pu,f, kN Pu ,bc, kN Pu ,bp, kN One-way shear Pu,1s, kN Two-way shear Pu,2s, kN Ultimate load Pu, kN
ACI 318-99 (1999) 902.5 3091.9 3496.8 864.1 890.1 864.1
ACI 318-14 (2014) 902.5 3091.9 3496.8 364.6 890.1 364.6
CRSI (2008) 902.5 3091.9 3496.8 1728.1 890.1 890.1

Notes: w is 125 mm (4.92 in.); d is 250 mm (9.84 in.); Pu,f = 2Mn,f /w, Mn,f is nominal flexural strength; Pu,bc = Pn,bc, Pn,bc is nominal bearing strength of column; Pu,bp = 4Pn,bp, Pn,bp
is nominal bearing strength of pile; Pu,1s = 2Vn,1s, Vn,1s is nominal shear strength of beam (ACI 318-99 Eq. (11.29), Eq. (a) of ACI 318-14 Table 22.5.5.1, Eq. (1)); Pu,2s = Vn,2s, Vn,2s
is nominal shear strength of two-way slab (Eq. (a) to (c) of ACI 318-14 Table 22.6.5.2, Eq. (2)); Pu = min(Pu,f, Pu,bc, Pu,bp, Pu,1s, Pu,2s); 1 kN = 0.2248 kip.

Table 2(b)—Evaluation of ultimate strength of pile cap BPC-30-25-1 by current design codes: ACI 318-14
strut-and-tie model approach
Concrete strut βs fc′, MPa fcu, MPa Fu, kN Aprov, mm2 Areq, mm2 Aprov /Areq Pu, kN
S1 0.51 29.1 14.84 321.5 13,717 21,663 0.63
Steel tie βt fy, MPa fcu, MPa Fu, kN Aprov, mm 2
Areq, mm 2
Aprov /Areq
T1 1.00 405 405 169.3 285 418 0.68
Nodal zone βn fc′, MPa fcu, MPa Fu, kN Aprov, mm2 Areq, mm2 Aprov /Areq
P/4 135.8 15,625 5491 2.85 543.3
CCC 0.85 29.1 24.74
S1 203.6 13,717 8230 1.67
R 135.8 17,671 9152 1.93
CTT 0.51 29.1 14.84 S1 203.6 22,963 13,717 1.67
T1 107.2 15,000 7225 2.08

Notes: βs, βt, and βn are coefficients of effective strength of strut, tie, and node; fcu is effective strength (= βs fc′ for strut, = βtfy for tie, = βnfc′ for node); FU is cross-sectional force
under experimental failure load; Aprov is maximum available area (refer to Fig. 2(b)); Areq is required area (= Fu/fcu); Pu is ultimate load (= minimum of Aprov/Areq × experimental
failure load); 1 kN = 0.2248 kip; 1 mm2 = 0.00155 in.2; 1 MPa = 145 psi.

d shown in Fig. 2(a) was used in this study. The load-carrying


Vc = Vc ≤ 0.83 f c′bd for w/d < 1.0 (1b) capacities of the elements were examined by comparing the
w ACI
maximum available cross-sectional areas of the elements
where fc′ is the compressive strength of concrete, in MPa; with the required cross-sectional areas. The available areas
w is the distance from the face of column to the center of concrete struts and nodal zones of the strut-and-tie model
of the nearest pile; b is the width of compression face of were determined using the ACI 445 (Klein 2002) approach,
the member; Vc ACI in Eq. (1a) is the same as Eq. (a) of which considers the geometrical shape of the selected strut-
ACI 318-14, Table 22.5.5.1; and Vc ACI in Eq. (1b) is the and-tie model and the sizes of the loading and bearing plates.
same as ACI 318-99 Eq. (11.29). For the shear design of a The available areas of the concrete struts and nodal zones
pile cap that can be regarded as a two-way slab, the section are shown in Fig. 2(b). The total cross-sectional area of the
at a distance d/2 from the face of column is considered to reinforcing bars placed within the effective width of steel
be the critical section and the nominal shear strength Vc tie T1 was taken as the maximum available area of the steel
at the critical section is determined from Eq. (a) to (c) of tie. The required areas of the struts and ties were determined
ACI 318-14, Table 2.6.5.2 and Eq. (2) for w/d ≥ 0.5 and w/d by dividing their cross-sectional forces by their effective
< 0.5, respectively strengths. The required areas of nodal zone boundaries were
determined by dividing the cross-sectional forces of the struts
d  d and ties framing the nodal zone by the effective strength of
Vc = 1 +  × 0.17 f c′bs d ≤ 2.67 f c′bo d (2)
w c the nodal zone. The procedure for evaluating the pile cap’s
ultimate strength is shown in Table 2(b) with an example of
where bo (mm) is the perimeter of critical section; and
the specimen BPC-30-25-1, wherein the minimum ratio of
bs equals 4 × c for a square column of dimension c. The
the maximum available area to the required area was taken
procedure for evaluating the ultimate strength of the pile
to determine its ultimate strength.
cap is shown in Table 2(a) with an example using specimen
The ultimate strength of the pile caps using the 3-D grid
BPC-30-25-1. The ultimate strengths of the other 77 pile
strut-and-tie model approach was determined by the linear
caps were determined in the same way.
analysis procedure presented in Yun et al. (2017). In the
The ultimate strength using the ACI 318-14 strut-and-tie
linear analysis procedure, the ultimate strength of a concrete
model approach was determined by examining the load-
member is obtained by estimating the maximum load that a
carrying capacities of the concrete struts, steel and concrete
3-D grid strut-and-tie model for the concrete member can
ties, and nodal zones of a selected strut-and-tie model. For the
carry by satisfying the condition of the strut-and-tie model’s
pile caps, the ACI 445 3-D strut-and-tie model (Klein 2002)

ACI Structural Journal/January 2018 29


Fig. 2—ACI 445 3-D strut-and-tie model for Pile Cap Fig. 3—3-D grid strut-and-tie model for Pile Cap BPC-30-
BPC-30-25-1. (Note: 1 kN = 0.2248 kip; 1 mm = 0.0394 in.; 25-1. (Note: 1 kN = 0.2248 kip; 1 mm = 0.0394 in.)
1 mm2 = 0.00155 in.2; 1 MPa = 145 psi.)
placed at the FE nodes located at the interfaces of the piles
geometrical compatibility and by evaluating the tensile and the pile cap, and the y-directional vertical roller was
stress of steel tie and the compressive stress of nodal zone placed at the FE node located at the bottom center of the pile
boundary at the maximum load. For the strength analysis of cap. A FE linear elastic analysis of the model was conducted
Specimen BPC-30-25-1, the 3-D finite element (FE) model and compressive principal stress flows were obtained. By
shown in Fig. 3(a) was constructed. The model is composed considering the stress flows, the position of the flexural
of eight-node unreinforced concrete elements. The external reinforcing bars, the clear cover, the locations of piles, and the
load acting on the column was distributed to the nodes of the ratio of vertical to horizontal length of a grid element in two
finite elements of the column by considering tributary areas directions, a 3-D grid strut-and-tie model composed of four
of the nodes. The x- and z-directional horizontal rollers were basic grid elements was constructed as shown in Fig. 3(b).

30 ACI Structural Journal/January 2018


Table 3(a)—Evaluation of ultimate strength of Pile Cap BPC-30-25-1 by 3-D grid strut-and-tie model
approach: ratio of cross-sectional areas under maximum load Pmax (=1158 kN)
Elements βs (βt) fc′, MPa fcu, MPa Aprov, mm2 Fu, kN Areq, mm2 Aprov /Areq
S1 1.15 29.1 33.57 16146 542.0 16146 1.000
T1 0.12 29.1 3.36 7784 25.4 7550 1.031
Notes: S1,T1: refer to Fig. 3(b); fcu = βs (βt) × fc′; Aprov is maximum available area in 3-D grid strut-tie model; Fu is cross-sectional force under maximum load Pmax; Areq is
required area (= Fu/fcu); 1 kN = 0.2248 kip; 1 mm2 = 0.00155 in.2; 1 MPa = 145 psi.

Table 3(b)—Evaluation of ultimate strength of Pile Cap BPC-30-25-1 by 3-D grid strut-and-tie model
approach: strength verification of steel tie under Pmax
Elements βt fy, MPa fcu, MPa Aprov, mm2 Fu, kN fs, MPa fy/fs Pfail, kN
T1 1.00 405 405 285.2 132.5 464.6 0.872 1009.7
Notes: fcu = βtfy; Fu is cross-sectional force under maximum load Pmax; fs = Fu/Aprov; Pfail = Pmax × fy/fs; 1 kN = 0.2248 kip; 1 mm = 0.00155 in. ; 1 MPa = 145 psi.
2 2

Table 3(c)—Evaluation of ultimate strength of Pile Cap BPC-30-25-1 by 3-D grid strut-and-tie model
approach: strength verification of nodal zone under Pmax
Connected
Nodal zone βn fc′, MPa fnz, MPa element σnz, MPa fnz/σnz Pfail, kN
CCC 3.45 29.1 100.5 S1 45.5 2.208 2559.2
CTT 0.60 29.1 17.5 S1 22.3 0.784 907.9

Notes: CCC, CTT: refer to Fig. 2(a); fnz = βnfc′; σnz is compressive stress of nodal zone boundary at Pmax; Pfail = Pmax × fnz/σnz; 1 kN = 0.2248 kip; 1 MPa = 145 psi.

The horizontal ties were placed at the center of the flexural and-tie model was greater than the yield strength fy (405
reinforcing bars. The external load acting on the column was MPa [58.7 ksi]) of the reinforcing bar with a ratio of fy/fs
distributed to the nodes (of the grid elements) located at the = 0.872, and the compressive stress σnz of the nodal zone
interface of the column and pile cap by considering tributary boundary framed by the concrete strut S1 at the CTT node
areas of the nodes. In the 3-D grid strut-and-tie model, the was greater than the effective strength fnz of the nodal zone
x- and z-directional horizontal rollers were placed at the FE with a ratio of fnz/σnz = 0.784. Therefore, according to the
nodes located at the centers of the piles. The y-directional definitions of the ultimate state of the strut-and-tie model’s
vertical roller was placed at the FE node located at the bottom components given in Yun et al. (2017), 907.9 kN (204.1 kip)
center of the pile cap. Following the procedure presented in (1158 kN [260.3 kip] × fnz/σnz, 105.8% of the experimental
Yun et al. (2017), the effective strengths of concrete struts failure load) was determined to be the ultimate strength of
were determined. The yield strength of the reinforcing bar Specimen BPC-30-25-1. The procedure for evaluating the
and the tensile strength of the concrete were taken as the ultimate strength of the specimen is shown in Tables 3(a),
effective strengths of the steel and concrete ties, respectively. 3(b), and 3(c). The ultimate strengths of the other pile caps
The required cross-sectional areas and the axial stiffness of were determined in the same way.
the struts and ties under the external load were determined The calculated ultimate strengths, coefficients of vari-
by using the simple iterative technique presented in Yun et ance, dispersions, and correlation coefficients of 78 pile
al. (2017). A structural analysis of the 3-D grid strut-and-tie caps are compared for several approaches in Fig. 4. As
model was conducted and the effective strengths of the shown in Fig. 4, the ultimate strengths were evaluated quite
nodal zones were determined by incorporating the structural conservatively with respect to test values using the ACI 318
analysis results. After carrying out structural analyses of sectional design method with the provisions of flexure,
the 3-D grid strut-and-tie model a few times according to bearing strength, and one-way slab shear. This indicates
the linear analysis procedure presented in Yun et al. (2017), that the use of a one-way slab shear provision is inappro-
the maximum load that the grid strut-and-tie model can priate for pile caps with w/d ≤ 5. The average ratios of the
resist by satisfying the conditions of nodal zone strength experimental failure strength to the calculated strength (and
and the strut-and-tie model’s geometrical compatibility the percentages of failure mode agreement) based on three
was determined. The dimensioned grid strut-and-tie model approaches: 1) ACI 318 provisions for flexure, bearing
under the maximum load is shown in Fig. 3(c). In the figure, strength, and two-way slab shear; 2) ACI 318 provisions
the cross-sectional areas of struts and ties are depicted as for flexure, bearing strength, and deep beam and two-way
circular shapes for the purpose of easier visual verification slab shear; and 3) the CRSI sectional design method; were
of the strut-and-tie model’s geometrical compatibility. The 1.13 (92.3%), 1.17 (69.2%), and 1.12 (92.3%), respectively.
maximum load that Specimen BPC-30-25-1 could resist The calculated strengths of more than 40% of the pile caps
was 1158 kN (260.3 kip) (135% of the experimental failure resulted in values greater than the experimental capacities
load). At the maximum load, the stress fs (464.6 MPa [67.4 by the three approaches. The average ratio of the experi-
ksi]) of the steel tie located at the bottom of the grid strut- mental strength to calculated strength (and the percentage of

ACI Structural Journal/January 2018 31


Fig. 4—Ultimate strengths of pile caps. (Note: 1 kN = 0.2248 kip.)
failure mode agreement) using the ACI 318-14 strut-and-tie ratios, patterns of shear reinforcing bars, and the shear
model approach was 2.29 (85.9%); therefore, resulting in reinforcement ratio are estimated using the 3-D grid strut-
excessive conservatism with respect to the experimental and-tie model approach and the two-way slab design
failure strength. However, the 3-D grid strut-and-tie model provisions of BS 8110 (1997), EC2 (2004), fib (2010), and
approach estimated the experimental failure strengths with ACI 318-14 (2014). Details of the slab-column joint tests are
reasonable accuracy when compared with the experimental summarized in Table 4. The geometry of the specimens and
capacity. The test to calculated capacity average ratio detailing of the flexural reinforcing bars are shown in Fig. 5
(and the percentage of failure mode agreement) was 1.12 and 6. All the test specimens were failed by punching shear.
(91.0%), overcoming the significant conservatism observed The ultimate strengths of the slab-column joints based on
with strut-and-tie model approaches and sectional design BS 8110 sectional design method were determined using
methods in the United States. the equation given in Table 3.8 of BS 8110 for the nominal
shear strength of a two-way slab. The ultimate strengths
ANALYSIS OF SLAB-COLUMN JOINTS TESTS of the slab-column joints based on EC2 and fib sectional
The ultimate strengths of 19 slab-column joints, tested by design methods were determined using EC2 Eq. (6.47), EC2
Yamada et al. (1992) and Ghannoum (1998) to investigate Eq. (6.52), and fib Eq. (7.3.40) and (7.3.42) for the nominal
the effects of tensile and compressive flexural reinforcement shear strength of a two-way slab. The ultimate strengths of

32 ACI Structural Journal/January 2018


Table 4—Specification of slab-column joints
No. of
Investigators Specimens specimens Slab size, mm Column size, mm fc′, MPa d, mm ρf ρv Ptest, kN
T-series 6 2000 x 2000 300 x 300 21.6 to 24.4 180 1.23 0.0 to 1.53 441 to 762
Yamada et al. (1992)
K-series 7 2000 x 2000 300 x 300 25.9 to 27.8 180 1.53 0.0 to 1.98 658 to 1498
U-type 3 2300 x 2300 225 x 225 37.2 to 67.1 125 0.96 — 301 to 443
Ghannoum (1998)
B-type 3 2300 x 2300 225 x 225 37.2 to 67.1 125 1.92 — 317 to 485

Notes: d is effective depth of slab; ρf, ρv are flexural and shear reinforcement ratios; Ptest is experimental failure load; 1 kN = 0.2248 kip; 1 mm = 0.0394 in.; 1 MPa = 145 psi.

Fig. 5—Geometries and details of flexural reinforcing bars of slab-column joints (Yamada et al. 1992). (Note: 1 mm = 0.0394 in.)
the slab-column joints using the ACI 318-14 sectional design model was conducted and the compressive principal stress
method were determined using Eq. (a) to (c) of ACI 318-14 flows were obtained. A 3-D grid strut-and-tie model was
Table 22.6.5.2 and Eq. (a) of ACI 318-14 Table 22.6.6.2. constructed by considering the compressive principal stress
The ultimate strengths of the slab-column joints using flows, the position of the flexural reinforcing bars, the clear
the 3-D grid strut-and-tie model approach were deter- cover and the ratios of the vertical-to-horizontal lengths of
mined using the linear analysis procedure shown in Yun a grid element in two directions, as shown in Fig. 7(b). The
et al. (2017). A detailed procedure for evaluating the ulti- horizontal tensile components of the grid elements were
mate strengths is illustrated using Ghannoum’s (1998) placed at the center of flexural reinforcing bars. The external
S1-U specimen. For the strength analysis of the specimen, loads acting on the loading plates were distributed to the FE
a 3-D FE model composed of eight-node unreinforced nodes (of the grid elements) located at the loading plates by
concrete elements was constructed, as shown in Fig. 7(a). considering the tributary areas of the nodes. As boundary
The external loads acting on the loading plates were distrib- conditions, the y-directional vertical roller at the FE node
uted to the nodes of the finite elements located at the loading located at the bottom center of the column and the x- and
plates by considering tributary areas of the nodes. A y- z-directional horizontal rollers at the other nodes located at
directional vertical roller was placed at the FE node located the bottom of the column, were placed. Following the proce-
at the bottom center of the column, and x- and z-directional dure presented in Yun et al. (2017), the effective strengths
horizontal rollers were placed at the other nodes located at of concrete struts were determined by reflecting the state
the bottom of the column. A FE linear elastic analysis of the of stresses of the 3-D FE model shown in Fig. 7(a). The

ACI Structural Journal/January 2018 33


Fig. 6—Geometry and detailing of flexural reinforcing bars of slab-column joints (Ghannoum 1998). (Note: 1 mm = 0.0394 in.)
yield strength of reinforcing bar and the tensile strength of
concrete were taken as the effective strengths of the steel
and concrete ties, respectively. The required cross-sectional
areas and axial stiffness of the struts and ties under the
external loads were determined by using the simple iterative
technique presented in Yun et al. (2017). After analyzing the
3-D grid strut-and-tie model a few times, the maximum load
that the grid strut-and-tie model can resist by satisfying the
strength conditions of the concrete struts was determined to
be 292 kN (65.6 kip) (97.0% of experimental failure load). At
the maximum load, the required cross-sectional areas of the
diagonal plane struts connected to the column in the bottom
x-z plane reached their maximum cross-sectional areas. The
dimensioned grid strut-and-tie model under the maximum
load is shown in Fig. 7(c). At the maximum load, the stress
at the four nodal zones located symmetrically at the bottom
of the grid strut-and-tie model, σcn (= 97.5 MPa [14.1 ksi]),
was greater than the effective strength of the nodal zones,
fcn (2.49fc′ = 92.2 MPa [13.4 ksi]). Therefore, according to
the linear analysis procedure of the approach with the defi-
nitions of ultimate states of strut-and-tie model components
(these are explained in Table 1 of Yun et al. 2017), 277.5 kN
(62.4 kip) (292 kN [65.6 kip] × fcn/σcn, 92.2% of the exper-
imental failure load) was determined to be the ultimate
strength of Specimen S1-U. Similarly, the ultimate strength
of the other slab-column joints was determined.
The average ratios of the experimental failure strength
to the evaluated strength based on the sectional design
methods of BS 8110, EC2, fib, and ACI 318-14 are 1.24,
0.98, 1.46, and 1.72, respectively, and the coefficients of
variation (COV) of each method are 35.8%, 35.2%, 34.3%,
and 52.9%, respectively. The 3-D grid strut-and-tie model
approach estimated the experimental failure strengths fairly
well and consistently with an experimental failure-to-
evaluated strength ratio and coefficient of variation of 1.16 Fig. 7—3-D grid strut-and-tie model for Slab-Column Joint
and 23.0%, respectively. All the test specimens by the 3-D S1-U. (Note: 1 mm = 0.0394 in.; 1 kN = 0.2248 kip.)

34 ACI Structural Journal/January 2018


Table 5—Specification of torsional beams
Investigators Specimens (A) fc′, MPa fly, MPa fty, MPa ρl ρt s, mm Ttest, kN-m
Series B 10 26.0 to 29.0 313.7 to 334.4 318.5 to 342.7 0.534 to 2.67 0.537 to 2.61 57 to 181 22.3 to 61.7
Series M 6 26.5 to 30.5 317.8 to 335.1 326.1 to 357.2 0.827 to 3.16 0.549 to 2.13 70 to 149 30.4 to 60.1
Series I 5 44.7 to 45.8 310.2 to 343.4 325.4 to 348.9 0.827 to 2.67 0.832 to 2.61 57 to 127 36.0 to 76.7
Hsu (1968)
Series J 4 14.3 to 16.9 319.9 to 338.5 331.6 to 346.1 0.534 to 1.60 0.537 to 1.61 92 to 152 21.5 to 40.7
Series G 8 26.8 to 31.0 319.2 to 338.5 321.3 to 349.6 0.400 to 1.58 0.402 to 1.60 86 to 187 26.8 to 73.5
Series N 6 27.3 to 30.4 330.9 to 352.3 337.8 to 360.6 0.611 to 1.42 0.622 to 1.42 51 to 92 9.0 to 15.7
Series B30 3 36.3 to 41.7 604.7 to 637.8 665.3 to 672.2 3.521 1.5 90 15.3 to 16.6

Rasmussen and Series B50 3 57.1 to 61.8 612.3 to 614.3 665.3 3.521 1.5 90 18.5 to 20.0
Baker (1995) Series B70 3 76.2 to 77.3 614.3 to 617.1 655.7 to 663.3 3.521 1.5 90 20.1 to 21.0
Series B110 3 105.0 to 109.8 617.8 to 634.3 655.0 to 659.8 3.521 1.5 90 23.6 to 24.8
Koutchoukali and
Series B 9 39.6 to 93.9 373.0 to 386.1 373.0 to 399.2 0.500 to 0.625 0.009 108 18.4 to 24.0
Belarbi (2001)

Notes: A is number of specimens; fly, fty are yield strengths of longitudinal and transverse reinforcing bars; ρl, ρt are longitudinal and transverse reinforcement ratios; s is spacing of
transverse reinforcing bars; Ttest is experimental failure load; 1 kN-m = 8.85 kip-in.; 1 mm = 0.0394 in.; 1 MPa = 145 psi.

Table 6—Ultimate strengths of torsional beams


No. of Rahal and Collins AASHTO-LRFD
Investigators specimens Hsu and Mo (1983) (1995) (2010) ACI 318-14 (2014) Present approach
Hsu (1968) 39 0.98 1.00 0.66 1.02 1.08
Rasmussen and
12 0.55 0.59 0.34 0.64 1.05
Baker (1995)
Koutchoukali and
9 0.82 0.80 0.67 1.01 1.02
Belarbi (2001)
Mean of Ttest/Tpredicted 0.87 0.88 0.60 0.94 1.06
Coefficient of variation, % 21.1 22.7 30.2 27.3 20.4

grid strut-and-tie model approach were failed by punching 2 Ao At f yt


shear (failure of inclined concrete struts encasing columns). Tn = cot θ (3)
s
This indicates that the approach can be used effectively for
the strength analysis and design of slab-column joints by where Ao is the gross area enclosed by the shear flow path
overcoming the inaccuracies and inconsistencies of current (≈0.85Aoh , where Aoh is the area enclosed by the centerline
sectional design methods. of the outermost closed transverse torsional reinforcement);
At is the area of one leg of a closed stirrup resisting torsion
ANALYSIS OF TORSIONAL BEAMS within spacing s; fyt is the yield strength of the transverse
The verification of the 3-D grid strut-and-tie model reinforcing bar; and θ is the angle between the axis of the
approach was carried out using the ultimate strength of 60 strut and the tension chord of the torsional beam. As the
reinforced concrete beams tested to pure torsion to failure by specimens tested were not prestressed, 45 degrees was taken
Hsu (1968), Rasmussen and Baker (1995), and Koutchou- as angle θ in applying the provisions of ACI 318-14. In
kali and Belarbi (2001). For the purposes of comparison applying the provisions of AASHTO-LRFD, the angle was
with current codes of practice, the ultimate strength of the determined by the method described in Section 5.8.3 of the
beams was also compared with AASHTO-LRFD (2010) and LRFD Specifications, incorporating the nominal shear stress
ACI 318-14 (2014) design provisions, Hsu and Mo (1983) and the longitudinal strain of the torsional beam. The details
method, and Rahal and Collins (1995) method. The data for of the methods by Hsu and Mo (1983) and Rahal and Collins
the torsional beams are listed in Table 5, and the geometry (1995) can be found in the references provided and is not
and details of the longitudinal reinforcement are shown in discussed herein.
Fig. 8. All the torsional beams were failed by yielding of The ultimate strength of each of the specimens tested was
transverse reinforcing bar(s) and diagonal cracks. determined using the nonlinear analysis procedure of the
The ultimate strength of the torsional beams based on the 3-D grid strut-and-tie model approach. The detailed proce-
design provisions of ACI 318-14 and AASHTO-LRFD was dure for evaluating the ultimate strength is illustrated with
determined using the following equation for the nominal an example of Beam B110-1 tested by Rasmussen and Baker
torsional strength Tn (kN) (1995). For the strength analysis of the beam, a 3-D FE
model for one-half of the effective span was developed, as

ACI Structural Journal/January 2018 35


Fig. 8—Geometries and details of flexural reinforcing bars of torsional beams. (Note: 1 mm = 0.0394 in.)

Fig. 9—3-D grid strut-and-tie model for Torsional Beam B110-1. (Note: 1 mm = 0.0394 in.)
shown in Fig. 9(a). The torsional force was linearly distrib- forcing bars, the clear cover of the concrete, and the ratios
uted to the nodes of the finite elements modeling the end of the vertical-to-horizontal lengths of a grid element in two
region of the beam. 3-D hinges were imposed at the nodes directions. The torsional force was distributed to the nodes
located at the mid-section of the beam. A FE linear elastic (of the basic grid elements) located at the end region of the
analysis of the model was performed, and the compressive beam by considering tributary areas of the nodes. 3-D hinges
principal stress flows were obtained as shown in Fig. 9(b). were imposed at the nodes of the grid elements located at the
As shown in Fig. 9(c), a 3-D grid strut-and-tie model was midsection of the beam. Following the procedure presented
constructed by considering the compressive principal stress in Yun et al. (2017), the effective strengths of the concrete
flows, the positions of transverse and longitudinal rein- struts were determined. The yield strength of reinforcing bar

36 ACI Structural Journal/January 2018


Fig. 10—Geometry, material properties, and load cases of pier cap. (Note: 1 mm = 0.0394 in.; 1 MPa = 145 psi.)
and the tensile strength of concrete were taken as the effec- proposed approach can be conservatively used for members
tive strengths of the steel and concrete ties, respectively. The under torsion. The failure of all the beams predicted by the
required cross-sectional areas and axial stiffness of the struts proposed approach occurred by diagonal cracks (failure of
and ties under the external loads were determined by using inclined concrete struts).
the simple iterative technique presented in Yun et al. (2017).
After analyzing the 3-D grid strut-and-tie model a few times, DESIGN OF PIER CAP
the maximum torsional force that the grid strut-and-tie model To verify the capability and effectiveness of the 3-D grid
can transfer by satisfying the strength conditions of concrete strut-and-tie model approach in the design of a concrete
struts was determined to be 95.2 kN-m (843.2 kip-in.) (385% member with multiple load combinations, we selected a
of experimental failure force). At the maximum torsional bridge pier cap subjected to three load cases with 3-D loads.
force, the required cross-sectional areas of the inclined The shear span-to-effective depth ratio of the cap is 1.25.
plane struts (Struts 742 and 743 in Fig. 9(c)) reached their The design conditions including geometry, material proper-
maximum cross-sectional areas. The dimensioned grid strut- ties, and load cases are given in Fig. 10.
and-tie model under the maximum torsional force is shown Following the linear analysis procedure presented in
in Fig. 9(d). To determine the ultimate strength of Spec- Yun et al. (2017), the design of the pier cap was conducted
imen B110-1 using the nonlinear analysis procedure of the using the 3-D grid strut-and-tie model shown in Fig. 11(a).
approach, a FE material nonlinear analysis of the 3-D grid The grid model was constructed using basic grid elements
strut-and-tie model was performed. The maximum torsional with vertical-to-horizontal length ratios of 0.96 and 1.09
force was applied incrementally with 40 load steps. At the in two directions. The auxiliary grid elements were also
10th step, the grid strut-and-tie model became unstable with used to model the inclined regions of the pier cap and the
sudden increases of nodal displacements. Therefore, by the outer regions of the round column. The AASHTO-LRFD
definitions of ultimate states given in Table 1 of Yun et al. (2010) strength-reduction factors of 0.75 and 0.90 were used
(2017), 23.8 kN-m (210.8 kip-in.) (95.2 × 10/40, 96.3% of for  the concrete struts and steel ties. The dimensioned strut-
experimental failure force) was determined to be the ulti- and-tie model for load case one is shown in Fig. 11(b). In Load
mate strength of Specimen B110-1. Following a similar Case 1 with the dead and live loads acting vertically, most of
approach, the ultimate strength of the other specimens tested the design loads at loading plates 2, 3, and 4 were transferred
was determined. to the pier directly by the concrete struts connecting the
The ultimate strength of each of the 60 torsional beams loading plates and column. On the other hand, most of the
evaluated with the existing and proposed methods are design loads at loading plates 1 and 5 were transferred to
compared in Table 6. The average ratios of the experimental the pier by the curved-shaped paths described by multiple
strength to the calculated strength by the methods or design concrete struts. The pier cap subjected to load cases 2 and
provisions of Hsu and Mo (1983), Rahal and Collins (1995), 3 with dead and live loads acting in three directions were
AASHTO-LRFD, and ACI 318-14 are 0.87, 0.88, 0.60, and designed with the same 3-D grid strut-and-tie model shown
0.94, respectively. It must be noted that all the methods over- in Fig. 11(b). This illustrates the versatility of the approach
estimated the experimental capacities. On the other hand, to handle multiple loading, including 3-D loads, by using a
the ultimate strengths were evaluated satisfactorily by the single type of 3-D grid strut-and-tie model. The conditions
proposed 3-D grid strut-and-tie model approach. Thus, the of the strut-and-tie model’s geometrical compatibility under

ACI Structural Journal/January 2018 37


which the column ends. The vertical positions of both nodes
are the same as those of corresponding nodes of the 3-D
grid strut-and-tie model shown in Fig. 11. The nodes were
also placed at a horizontal distance of 620 mm (24.4 in.)
from the boundary of the column with a transformed square
cross section. The horizontal positions of both nodes can be
moved toward the boundary of the pier by considering the
required cross-sectional areas of the concrete struts located
in the pier.
The required cross-sectional areas of reinforcing bars
obtained using each method are compared in Fig. 13. From
the design results, two features can be observed. First, more
flexural reinforcing steel is required by the strut-and-tie
model approach than by the other methods. This is mainly
due to the positions of nodes F and G shown in Fig. 12 and
the corresponding nodes shown in Fig. 11. If these nodes
are moved toward points F and G by considering the critical
section (the S-S line in Fig. 12) defined by ACI 318-14 flex-
ural design and bracket provisions, similar flexural design
results will be obtained. Second, more vertical shear rein-
forcing bars are required by the strut-and-tie model of truss
mechanism than by the 3-D grid strut-and-tie model. This
is because the design loads at loading plates 1 and 5 are
transferred to the column mainly by truss action in the strut-
and-tie model of the truss mechanism. On the other hand, the
design loads are transferred to the pier by the combined arch
and truss mechanisms in the 3-D grid strut-and-tie model.
In other words, a portion of the design loads are transferred
directly to the pier through arch-type concrete struts in the
3-D grid strut-and-tie model. These design outcomes of
the pier cap (with a shear span-to-effective depth ratio of
1.25) based on the 3-D grid strut-and-tie model approach
are supported by the statically indeterminate strut-and-tie
models recommended by Foster and Gilbert (1998), fib
Fig. 11—3-D grid strut-and-tie model for pier cap. (Note: (2010), and Kim and Yun (2011). The recommended models,
1 kN = 0.2248 kip.) consisting of the arch and truss load transfer mechanisms,
are known to be most appropriate for deep beams with shear
all three load cases were satisfied, and the strengths of all
span to effective depth ratios of 0.45 < a/d < 1.80, 0.90 < a/d
nodal zones were enough to transfer the strut and tie forces.
< 1.56, and 0.50 < a/d < 3.00, respectively.
Thus, the maximum cross-sectional area of a steel tie at a
certain position under the three load cases was taken as the
SUMMARY AND CONCLUSIONS
required area of the reinforcing bar that should be distributed
The strut-and-tie model approach in current design codes
with the effective area of the steel tie. The required areas
incorporated with simple truss-type 2-D strut-and-tie models
of reinforcing bars corresponding to the other steel ties
for the design of structural concrete with disturbed regions
were obtained and distributed in the same way. To complete
in most cases is not sufficient to properly represent the
the design, the code provisions for auxiliary or minimum
complex load transfer mechanisms encountered in some 3-D
reinforcement must be observed.
disturbed regions. To overcome this limitation, the authors
To compare the design results of the 3-D grid strut-and-tie
have proposed a 3-D grid strut-and-tie model approach in
model approach, the designs of the pier cap were carried out
the authors’ previous paper. The approach employs grid
using the ACI 318-14 flexural design method, ACI 318-14
strut-and-tie models to represent the complex load transfer
bracket provisions, and ACI 318-99 deep beam provisions.
mechanisms, and to determine the effective strengths of 3-D
The pier cap was also designed based on the strut-and-tie
concrete struts and nodal zones. The approach is envisioned
model approaches of current design codes, with the
for use in practice and not only as a research tool. The same
two-dimensional (2-D) strut-and-tie models shown in Fig.
basic concepts have been adopted of the strut-and-tie model
11(a) and 11(b) representing arch and truss load-transfer
approaches used in current design codes in terms of the
mechanisms, respectively. As the 2-D strut-and-tie models
strength verification of struts, ties, and nodal zones.
are valid for 2-D load cases, only Load Case 1 (Fig. 11)
The approach was used to estimate the experimental capac-
was considered. Nodes F and G in Fig. 12 were placed at
ities of pile caps, slab-column joints, and torsional beams.
a vertical distance of 200 mm (7.87 in.) from the plane in
The average ratios of the experimental strength to calculated

38 ACI Structural Journal/January 2018


Fig. 12—Two-dimensional strut-and-tie model for pier cap. (Note: 1 kN = 0.2248 kip; 1 mm = 0.0394 in.)

Fig. 13—Design results of pier cap. (Note: 1 cm2 = 0.155 in.2)

ACI Structural Journal/January 2018 39


strength (and COV) by ACI 318-14 sectional design methods ACI Committee 318, 1999, “Building Code Requirements for Struc-
tural Concrete (ACI 318-99) and Commentary (ACI 318R-99),” American
for pile caps, slab-column joints, and torsional beams were Concrete Institute, Farmington Hills, MI, 369 pp.
1.13 (29.3%), 1.72 (52.9%), and 0.94 (27.3%), respectively. ACI Committee 318, 2014, “Building Code Requirements for Struc-
The average ratios (and COV) by EC 2 and fib for slab- tural Concrete (ACI 318-14) and Commentary (ACI 318R-14),” American
Concrete Institute, Farmington Hills, MI, 519 pp.
column joints were 0.98 (35.2%) and 1.46 (34.3%). The British Standards Institution, 1997, “Structural Use of Concrete: BS
average ratios (and COV) by AASHTO-LRFD and Rahal 8110,” Milton Keynes, UK.
and Collins (1995) method for torsional beams were 0.60 Clarke, J. L., 1973, “Behavior and Design of Pile Caps with Four Piles,”
Report No. 42.489, Cement and Concrete Association, London, UK.
(30.2%) and 0.88 (22.7%). The average ratios (and COV) by CEB-FIP, 2010, “CEP-FIP Model Code,” Comité Euro-International
the proposed approach for pile caps, slab-column joints, and du Béton, International Federation for Structural Concrete (fib), Lausanne,
torsional beams were 1.12 (12.8%), 1.16 (23.0%), and 1.06 Switzerland.
Concrete Reinforcing Steel Institute, 2008, CRSI Handbook, Chicago, IL.
(20.4%), respectively. The approach evaluated the experi- CEN, 2004, “Eurocode 2: Design of Concrete Structures,” European
mental capacities fairly well and consistently, and enabled Committee for Standardization, Brussels, Belgium, 229 pp.
the design of the pier cap by representing its load transfer Foster, S. J., and Gilbert, R. I., 1998, “Experimental Studies on High-
Strength Concrete Deep Beams,” ACI Structural Journal, V. 95, No. 4,
mechanisms properly. By considering the strength analysis July-Aug., pp. 382-390.
and design results of this study, the 3-D grid strut-and-tie Ghannoum, G. M., 1998, “Effect of High-Strength Concrete on the
model approach can help overcome the limitations and Performance of Slab-Column Specimens,” MS thesis, Department of
Civil Engineering and Applied Mechanics, McGill University, Montreal,
uncertainties of the strut-and-tie model approaches used in QC, Canada.
current design codes. Hsu, T. T. C., 1968, “Torsion of Structural Concrete—Interaction Surface
for Combined Torsion, Shear, and Bending in Beams Without Stirrups,”
ACI Journal Proceedings, V. 65, No. 1, Jan., pp. 51-60.
AUTHOR BIOS Hsu, T. T. C., and Mo, Y. L., 1983, “Softening of Concrete in Torsional
Young Mook Yun is a Professor in the Department of Civil Engineering
Members,” Research Report No. ST-TH-001-83, Department of Civil Engi-
at Kyungpook National University, Daegu, Korea. He received his PhD
neering, University of Houston, Houston, TX.
from Purdue University, West Lafayette, IN, in 1995. His research inter-
Kim, B. H., and Yun, Y. M., 2011, “An Indeterminate Strut-Tie Model
ests include strut-and-tie model analysis and design of concrete members,
and Load Distribution Ratio for RC Deep Beams—(I) Model & Load
computational mechanics, and computer graphics.
Distribution Ratio,” Advances in Structural Engineering, V. 14, No. 6,
pp. 1031-1042.
Hyun Soo Chae is a Deputy Head in Program Development Department
Klein, G. J., 2002, “Example 9: Pile Cap,” Examples for the Design for
of Hangil IT, Seoul, Korea. He received his PhD from Kyungpook National
Structural Concrete with Strut-and-Tie Models, SP-208, K.-H. Reineck, ed.,
University in 2012. His research interests include numerical and strut-
American Concrete Institute, Farmington Hills, MI, pp. 213-223.
and-tie model analysis/design of disturbed regions in concrete members.
Koutchoukali, N., and Belarbi, A., 2001, “Torsion of High-Strength
Reinforced Concrete Beams and Minimum Reinforcement Requirement,”
Byunghun Kim is a Lead Engineer in the Infrastructure Department of
ACI Structural Journal, V. 98, No. 4, July-Aug., pp. 462-469.
Hyundai Engineering, Seoul, Korea. He received his PhD from Kyungpook
Rahal, K. N., and Collins, M. P., 1995, “Analysis of Sections Subjected
National University in 2004. His research interests include strut-and-tie
to Combined Shear and Torsion—A Theoretical Model,” ACI Structural
model design of concrete structures and development of numerical analysis
Journal, V. 92, No. 4, July-Aug., pp. 459-469.
methods for steel and reinforced concrete structures.
Rasmussen, L. J., and Baker, G., 1995, “Torsion in Reinforced Normal
and High-Strength Concrete Beams—Part 1: Experimental Test Series,”
Julio A. Ramirez, FACI, is a Professor in the Lyles School of Civil Engi-
ACI Structural Journal, V. 92, No. 1, Jan.-Feb., pp. 56-62.
neering at Purdue University. He is a member of Joint ACI-ASCE Commit-
Sabnis, G. M., and Gogate, A. B., 1984, “Investigation of Thick Slab
tees 408, Bond and Development of Steel Reinforcement, and 445, Shear
(Pile Cap) Behavior,” ACI Journal Proceedings, V. 81, No. 1., Jan.-Feb.,
and Torsion. He has received the 2000 ACI Delmar Bloem Award and the
pp. 35-39.
2006 ACI Joe W. Kelly Award.
Suzuki, K.; Otsuki, K.; and Tsubata, T., 1998, “Influence of Bar Arrange-
ment on Ultimate Strength of Four-Pile Caps,” Transactions of the Japan
ACKNOWLEDGMENTS Concrete Institute, V. 20, pp. 195-202.
This research was supported by Basic Science Research Program through Suzuki, K.; Otsuki, K.; and Tsuchiya, T., 2000, “Influence of Edge
the National Research Foundation of Korea (NRF) funded by the Ministry Distance on Failure Mechanism of Pile Caps,” Transactions of the Japan
of Education, Science and Technology (NRF-2015R1D1A1A01061333). Concrete Institute, V. 22, pp. 361-368.
Yamada, T.; Nanni, A.; and Endo, K., 1992, “Punching Shear Resistance
of Flat Slabs: Influence of Reinforcement Type and Ratio,” ACI Structural
REFERENCES Journal, V. 88, No. 4, July-Aug., pp. 555-563.
American Association of State Highway and Transportation Officials, Yun, Y. M.; Kim, B. H.; and Ramirez, J. A., 2017, “Three-Dimensional-
2010, AASHTO LRFD Bridge Design Specifications, fifth edition, Wash- Grid Strut-and-Tie Model Approach in Structural Concrete Design,” ACI
ington, DC. Structural Journal, V. 115, No. 1, Jan., pp. 15-26.

40 ACI Structural Journal/January 2018


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 115-S03

Time-Dependent Buckling Testing of Eccentrically Loaded


Slender High-Strength Concrete Panels
by Yue Huang, Ehab Hamed, Zhen-Tian Chang, and Stephen J. Foster

The results of nine full-scale high-strength concrete panels tested consider the P-Δ effect. Therefore, there is a need to look
under eccentric sustained loads are presented in this paper. The at the combined geometric and material nonlinearity of the
panels were uniaxially loaded for up to 4 months. Four panels problem. However, in most cases, a considerable portion of
failed due to creep buckling at different times under sustained loads the load can be classified as a sustained load that leads to
that are lower than their short-term capacities. The other panels
an increase of the axial and out-of-plane deformations with
exhibited a long-term stable behavior; they were then loaded to
time due to creep and the geometric nonlinearity, which
failure without the release of the existing sustained loads and
exhibited a reduction in their residual strength due to creep. The increase the bending moments due to the increased P-Δ
study investigates the influences of aging of concrete, magnitude effect as a result of increasing Δ with time. This may lead to
of the applied load, and the slenderness ratio. The experimental cracking of the wall and, consequently, may lead to loss of
results are compared to theoretical predictions generated from a stability at certain time—so-called “creep buckling.” Such
theoretical analysis that was previously developed by the authors. behavior was observed in reinforced concrete (RC) columns
The formulation is based on the rheological generalized Maxwell and shells.1-3 The creep and shrinkage deformations may not
model and it accounts for large deformation kinematics in the necessarily lead to buckling failure, but they may increase
structural level. A close correlation is achieved between the exper- the internal stresses and decrease the residual strength and
imental and theoretical results. the factor of safety of the wall.
Keywords: buckling; cracking; creep; high-strength concrete; panels.
The instantaneous buckling behavior of NSC and HSC
panels has been extensively investigated both theoretically
INTRODUCTION and experimentally by numerous studies.4-7 Nevertheless,
The use of high-strength concrete (HSC), whose compres- no experimental study could be found in the literature
sive strength is over 55 MPa (7975 psi), has been rapidly regarding the influence of creep on the buckling capacity of
growing over the past decades due to its superior prop- HSC panels, which is the contribution of this paper. Among
erties such as higher strength, improved durability, and the studies that focused on the buckling response of panels
lighter-weight construction in comparison to conventional without creep, Oberlender and Everard4 tested 54 one-way,
normal-strength concrete (NSC). HSC panels have been simply supported NSC panels to failure with varying slen-
widely used in engineering structures such as load-bearing derness ratios, concrete strength, and reinforcement ratios.
walls, core walls in high-rise buildings, box girders in Saheb and Desayi5 studied the influence of vertical and hori-
bridges, and hulls of offshore structures. Usually, HSC zontal reinforcement ratios, as well as the slenderness and
panels are more slender than typical NSC panels because aspect ratio (height/width) on the behavior of one-way NSC
of their higher stress capacity. The increased slenderness wall panels. The ultimate strength of the panel was found to
of the panel highlights the need for investigating the buck- vary linearly with respect to the vertical reinforcement ratio
ling capacity of the panel, and its degradation with time due and to the aspect ratio. Fragomeni and Mendis6 numerically
to creep, shrinkage, and time-dependent cracking of the studied the stability of NSC and HSC wall panels. Huang
concrete. This is the focus of this study. et al.7 tested eight full-scale HSC panels under short-term
Vertical HSC wall panels are subjected to in-plane eccentric loading with different load eccentricities, reinforce-
compressive loads and out-of-plane bending moments. The ment ratios, and slenderness ratios. A theoretical model
moments can be a result of vertical load eccentricity, initial that considered material nonlinearity, cracking, and the
imperfection, construction inaccuracy, and moments due tension-stiffening effect was also developed and validated.
to dynamic lateral actions such as wind gusts and seismic While no experimental studies regarding the time-dependent
inputs. Such loading combinations may lead to buck- buckling behavior of wall panels could be found in the litera-
ling failures of slender panels or material failures in short ture, a number of experimental works have been undertaken
panels that are characterized by crushing of the concrete to study the geometrically nonlinear response (P-Δ effect)
and/or yielding of the steel reinforcement. In moderately of RC columns under sustained loads.8-13 Despite the simi-
slender walls, which is the most common case in practice,
such material failures can be triggered or accelerated by ACI Structural Journal, V. 115, No. 1, January 2018.
MS No. S-2015-382.R2, doi: 10.14359/51700913, received February 22, 2017, and
the geometric nonlinearity (P-Δ effect), and they can be reviewed under Institute publication policies. Copyright © 2018, American Concrete
observed at load levels that are much lower than what is Institute. All rights reserved, including the making of copies unless permission is
obtained from the copyright proprietors. Pertinent discussion including author’s
predicted from a geometrically linear analysis that does not closure, if any, will be published ten months from this journal’s date if the discussion
is received within four months of the paper’s print publication.

ACI Structural Journal/January 2018 41


larities in the modeling and behavior between RC columns
and one-way panels, there are some differences that require
attention and justify the need for creep testing of HSC
panels. The concrete in panels is unconfined due to the
absence of stirrups, and panels have significantly smaller
reinforcement ratios. Due to these differences, the role and
importance of the tension-stiffening effect is magnified, but
most importantly, there is less restraint to the creep defor-
mations of the concrete. Among the studies that focused on
the time-dependent creep buckling response of RC columns,
Green and Breen8 tested 10 identical NSC columns under
eccentric sustained loads. Two columns failed under the
sustained loads within 2 months. The remaining columns
continued to deform with time with no indication of failure
for 1.5 years. Goyal and Jackson9 reported experimental Fig. 1—Schematic description of tested panels and
results of 20 NSC columns tested under sustained loading. typical cross section. (Note: All dimensions in mm; 1 mm
The residual strength of the long-term loaded columns = 0.0394 in.)
ranged from 82% to 102% of the short-term capacity.
ison aims to validate the model, which can then be used to
Claeson and Gylltoft12 tested one NSC column and one HSC
generate comprehensive parametric studies.
column under eccentric sustained loading. The NSC column
failed after 22 days under a sustained load level of 80% of
RESEARCH SIGNIFICANCE
the short-term failure load, while the HSC column exhibited
Slender HSC wall panels may experience excessive out-of-
little increase in the deflection after 6 months and thereby
plane deflections and cracking over time under combined
was loaded to failure at 190 days. The load level at failure
axial and transverse loads due to creep and shrinkage. More-
was 85 to 89% of its accompanying column that was tested
over, they may fail by creep buckling. A significant number
on the same day under short-term loading, which highlighted
of studies have been conducted to study the instantaneous
the influence of creep on the load-carrying capacity.
behavior of RC panels and the time-dependent responses
A limited number of studies have been undertaken to
of RC columns. However, no studies have focused on the
investigate the time-dependent behavior of RC panels in
time-dependent buckling behavior of HSC panels, which are
the literature. However, they focused on the serviceability
more slender than typical NSC panels. In this study, nine HSC
response rather than the ultimate buckling capacity of
panels were tested under sustained axial loads, which showed
the panel under sustained loads. Moreover, the panels in
that creep can lead to a significant reduction in the load-car-
these studies were made of NSC rather than HSC. Tatsa14
rying capacity of HSC wall panels. The test results were also
tested seven concrete panels under low levels of eccentric
used to validate a theoretical model that was previously devel-
sustained load at one end and concentric sustained load at
oped by the authors,16 and which can be further used in the
the other end. It was observed that the neutral axis shifts
future for conducting more comprehensive parametric studies
toward the compression face with time. It was also found
that are beyond the specific geometry, materials, and loadings
that there was a 23% reduction in the residual strength of
used in this experimental program. The theoretical model of
the panel when tested for failure after 4 months of creep.
Huang and Hamed16 was selected because it has many advan-
Lee et al.15 investigated the time-dependent behavior of NSC
tages over existing models and it can describe the structural
panels, where the sustained load was applied concentrically
behavior at different scales. The experimental results also
on part of the entire width of the wall. Following the test
provide a benchmark database to the literature.
results, and in conjunction with a finite element analysis, an
effective width coefficient was proposed to be included in
EXPERIMENTAL INVESTIGATION
the structural analysis and design of concrete panels.
Test specimens
This paper reports the results of nine panels that were
The experimental investigation includes testing of nine
tested to investigate the long-term behavior of one-way HSC
slender, high-strength reinforced concrete wall panels under
panels under eccentric sustained in-plane loads. The panels
eccentric sustained in-plane compressive loads along their
were simply supported along the two short edges. The influ-
short edges. Panels LT1 to LT8 had the same dimensions,
ences of the loading age, level of the in-plane load, and slen-
where the length (L), width (b), and thickness (h) were 2700,
derness ratio on the time-dependent behavior were studied.
460, and 100 mm (106.3, 18.1, and 3.94 in.), respectively
The test findings represent a benchmark database that can be
(Fig. 1). Panel LT9 had the same length and width as LT1
used for further long-term stability studies of HSC panels.
to LT8 but with a thickness of 130 mm (5.12 in.). All panels
Moreover, the experimental results are compared with the
were equally reinforced at the top and bottom layers and
estimations generated by a theoretical model proposed by
had a total reinforcement ratio of 0.22% for LT1 to LT8
the first two authors (Huang and Hamed16). The theoret-
and 0.28% for LT9. Welded wire reinforcement SL52, with
ical model considers geometric nonlinearity in conjunction
bar diameter of 4.77 mm (0.188 in.) at spacing of 200 mm
with concrete creep, shrinkage, and cracking. The compar-
(7.87 in.) in two orthogonal directions, was used as the

42 ACI Structural Journal/January 2018


Fig. 2—Test setup.

Table 1—Tested panels


Concrete age upon Sustained load Predicted short-term edesign*, mm
Panel No. Batch No. L/r loading (t0), days level, kN failure load Pu, kN Sustained load/Pu [edesign/h] eL†, mm
2
eR‡, mm
3

LT1 1 94 22 591 638 93% 16.7 [1/6] 22.5 21.9


LT2 94 146 320 436 73% 33.3 [1/3] 31.1 37.5
2
LT3 94 99 615 820 75% 16.7 [1/6] 19.1 18.8
LT4 94 22 369 499 74% 25 [1/4] 29 31
3
LT5 94 22 548 764 72% 16.7 [1/6] 19.6 21.4
LT6 94 22 490 747 66% 16.7 [1/6] 21.7 20
4
LT7 94 22 808 1526 53% 8.3 [1/12] 7.2 7.1
LT8 94 22 927 1301 71% 8.3 [1/12] 9.5 8.6
5
LT9 72 22 1020 1075 95% 21.7 [1/6] 32.6 28.1
*
Designed eccentricity.

Measured eccentricity at left end of tested panel.

Measured eccentricity at right end of tested panel.
Notes: 1 mm = 0.0394 in.; 1 kN = 0.225 kip.

reinforcement for Panels LT1 to LT8, and SL72 mesh with HSC slender panels (Huang et al.7) has shown that the behavior
bar diameter of 6 mm (0.236 in.) and spacing of 200 mm is greatly influenced by the load eccentricity and slenderness
(7.87 in.) was used for Panel LT9. The concrete cover was of the panel, and much less influenced by the reinforcement
20 mm (0.787 in.) for all panels. The test setup is schematically ratio. Based on these findings and due to time limitations, it
depicted in Fig. 2, along with a typical 100 mm (3.94 in.) thick was decided not to investigate this parameter. The sustained
cross section. Although the panels are intended to work verti- load ratio for each panel is determined as the ratio of the
cally, they were tested horizontally for safety and simplicity measured sustained load divided by the corresponding short-
reasons. This is justified because the level of bending caused term failure load, which was predicted by a theoretical model
by the self-weight and any other imperfections is negligible developed in Huang et al.7 In Huang et al.,7 the model was
compared to the bending caused by the load eccentricity. compared with test results generated by the authors for panels
Three parameters were examined in the tests, including the possessing the same dimensions as the ones tested herein, and
age of loading; the ratio of the sustained load to the failure with other test results from the literature.5,17 For the predic-
load; and the effective slenderness ratio, defined as L/r, where tion of the failure loads, the measured material properties
r is the radius of gyration = r = I eff /Aeff , and Ieff and Aeff are and actual eccentricities were used. As the sustained loading
the effective moment of inertia and cross section area, respec- testing normally takes a significantly long time, the load ratios
tively. The details of the loading conditions are reported in for the specimens in the present study were designed to be
Table 1. Although the reinforcement ratio is a parameter that higher than conventional sustained loading tests so that the
needs to be considered in general, a previous study conducted panels might fail within a short time since loading (that is, up
by the authors regarding the short-term buckling response of to 6 months).

ACI Structural Journal/January 2018 43


Table 2—Material properties of concrete at start
and end of panel testing
Elastic Tensile
Batch Concrete Compressive modulus Ec, strength ft,
No. age, days strength fc′, MPa GPa MPa
22 80.9 39.4 5.3
1
38 89.5 40.3 6.4
99 91.7 40.25 6.9
2
203 95 40.3 7
22 101 39.5 5.9
3 91 103.5 39.8 8.9
126 103 40.5 8.9
22 89.1 39.5 6.2
4
134 100 43.4 7.7
22 88 39.4 5.5
5 30 90.5 39.8 6.0
104 98 41.1 6.5

Notes: 1 MPa = 0.145 ksi; 1 GPa = 145 ksi.

The eccentricities reported herein include the design


eccentricities (edesign) as well as the actual eccentricities at
the left (eL) and right edges (eR) of the panel obtained from
the test. The actual eccentricities were determined as follows

Ec I eff ε top − ε bot


etest = (1)
N test h

where etest is the actual eccentricity in the test at the left edge
(eL) or the right edge (eR) (refer to Table 1); Ntest, εtop, and εbot
are the experimental axial load, and the measured concrete
strain at the top and bottom faces of the panel, respectively.
The effective moment of inertia is obtained using the trans-
formed section method. Equation (1) is based on a linear Fig. 3—Details of test setup.
strain distribution through the thickness of the panel and Preparation of test specimens
a linear material behavior. As the out-of-plane deflections The formwork of the panel specimens was built using
near the panel ends are small, the geometric nonlinearity is structural-grade plywood and laid horizontally on the ground
neglected in Eq. (1). The mean values of the experimental in the laboratory. The SL52 mesh was cut to the required
eccentricities are used, which were determined at a short- dimension and placed at the top and bottom layers that were
term load level equal to 30% of the estimated failure load. held in place by steel bar chairs. The panels were cast using
Panels LT2 and LT3 were loaded at a concrete age of a commercial high-strength concrete. They were made from
146 and 99 days, respectively, whereas the other panels five concrete batches with the same concrete mixture design
were all loaded at the same concrete age of 22 days. The (refer to Tables 1 and 2). After casting, the panels were
testing age (22 days) was selected from the considerations covered with wet hessian and with plastic sheets over the
of having sufficient strength while allowing significant creep hessian. They were kept moist in the formwork for 14 days
so that some critical aspects of the structural behavior can be before being stripped; they then remained in the laboratory
observed or revealed within a reasonable time period. The at ambient conditions until the day of loading.
effect of loading age is investigated by comparing the test
results of Panel LT3 to those of LT5. The influence of the Test setup and instrumentation
in-plane load level is studied by comparing the results of The panels were tested in a testing frame with a hydraulic
Panels LT1, LT5, and LT6. The influence of the slenderness jack mounted at one end and two load cells mounted at the
ratio is examined by comparing the results of Panel LT1 and other end. The load was applied and monitored to remain
LT9. Panel LT4 was initially loaded to 93% of its predicted constant with time through the hydraulic jack. The test setup
short-term failure load, followed by an unloading stage until is shown in Fig. 2 and 3. Two groups of disc springs with
the load was dropped to 74% of the failure load, which was high load-carrying capacity were inserted between the jack
then kept constant with time. and the end support plate to minimize the reduction of the

44 ACI Structural Journal/January 2018


Fig. 4—Location of strain gauges. (Note: All dimensions in mm; 1 mm = 0.0394 in.)
axial sustained load with time due to creep and relaxation. humidity, but the temperature would typically be at room
The loading and supporting mechanisms at each panel end temperature, varying between 20 and 25°C (68 and 77°F),
consisted of a rotatable knife-hinge to provide a simply and the humidity would be varying between 40 and 60%.
supported boundary condition (refer to Fig. 3). The load
eccentricity was set up in a way that the panel would deform Material properties
upwards under loading. The out-of-plane displacements Concrete cylinders 100 mm (3.94 in.) in diameter and
were measured by laser displacement sensors at three points 200 mm (7.87 in.) high, and concrete prisms of 100 x 100
that were symmetrically positioned along the top face of the x 500 mm (3.94 x 3.94 x 19.7 in.) were cast and cured from
panel. Two sensors were located at 100 mm (3.94 in.) away each batch during the preparation of the panels. The control
from each edge, and one was placed at the center length. cylinders and prisms were kept in the same environment
Eighteen strain gauges were placed on the surfaces of steel as the panels and were similarly cured. The concrete cylin-
and concrete at critical locations as shown in Fig. 4. The ders were tested in compression to measure the compres-
displacements of the hinge plates at the two ends were moni- sive strength, the elastic modulus, and the full stress-strain
tored using a linear strain conversion transducer (LSCT), as behavior. The prisms were tested under four-point bending
shown in Fig. 3. to determine the flexural tensile strength. The concrete
mechanical properties were measured at the commencement
Test procedure and completion of each panel test. The material test results
Each panel was loaded up to the desired sustained load are reported in Table 2.
level and the load was then held constant during the test. Two concrete prisms of 75 x 75 x 280 mm (2.95 x 2.95 x
The duration of applying the loads was less than 30 minutes. 11 in.) with gauge studs mounted on both ends were cast and
The sustained load dropped as a result of concrete relax- cured with each batch in order to measure the shrinkage strain
ation. The dropping was remarkable, especially at the first in accordance with Australian Standard AS1012.13.18 The
few hours after the sustained load was applied. Therefore, prisms were stripped off at the same day as the panels and then
the load was topped up from time to time, as needed, so kept in the same environment as the panels. The measurement
that it stayed at a nearly constant level throughout the test of the shrinkage strain started immediately after the panels
to simulate creep behavior. The procedure of adjusting the were demolded at 14 days by using a vertical comparator, and
load was made using a hydraulic pump that was connected stopped after failure of the panel. The measured shrinkage
to the loading jack. If the tested panel was found to show data for each concrete batch is shown in Fig. 5.
very little increase in the deflection over a certain period Three cylinders from each batch were placed in a creep
after sustained loading, it was considered to have long-term test rig under a sustained stress of 30 MPa (4.35 ksi) when
stability and the test would be terminated. The panel was each group of panels that are made from the same batch were
then loaded to failure without releasing the existing sustained tested. The stress was applied hydraulically on the same day
load following the same procedure used for adjusting the of undertaking the long-term test of the corresponding panel,
load level. All tests were conducted under typical ambient and was adjusted to maintain constant over time. The total
indoor conditions within the laboratory. Unfortunately, no strains—which included the elastic strain, the shrinkage
measurements were made to monitor the temperature and and thermal strain, and the creep strain—were measured

ACI Structural Journal/January 2018 45


from the three cylinders using Demec strain gauges. The than the other batches because it was tested at older ages
shrinkage and thermal strains were measured from two upon testing Panels LT2 and LT3. The curves of the concrete
unloaded companion cylinders. The creep strains were made from Batches 1 and 5 have a similar trend, as so do the
then determined by subtracting the sum of the measured curves of concrete Batches 3 and 4. However, the creep coef-
shrinkage and thermal strain and instantaneous elastic strain ficients of Batches 3 and 4 appear to be smaller compared
from the total strains. The creep coefficient, defined as the to those of Batches 1 and 5, although all concrete cylinders
ratio of measured average creep strain to the measured were loaded at the same age of 22 days. The reason for that
average elastic strain, for each concrete batch is shown in could be mainly attributed to the fact that the batches were
Fig. 6. Batch 2 exhibited noticeably smaller creep over time tested at different seasons. Batches 1 and 5 were tested in
winter, whereas Batches 3 and 4 were tested around summer
with much higher relative humidity, which resulted in less
creep. The differences in the temperature between the two
seasons in indoor conditions are too small to significantly
influence the creep response.
Samples of the reinforcing steel were tested to determine
the stress-strain properties. The elastic modulus and yield
strain were measured as 206 GPa (29,878 ksi) and 2 × 10–3.
The measured compressive strains of concrete corresponding
to the compressive strength in the stress-strain curves ranged
from 3.5 × 10–3 to 3.8 × 10–3.

EXPERIMENTAL RESULTS AND DISCUSSIONS


Fig. 5—Shrinkage strains measured in test. Both the experimental results and the predictions by the
theoretical model of Huang and Hamed16 are summarized
in Table 3. In the model,16 the concrete is assumed to be
linear viscoelastic in compression and nonlinear in tension,
where cracking and tension stiffening are included following
Fields and Bischoff.19 The measured material properties
are incorporated into the long-term theoretical model. The
aging effect of concrete, which includes time variation of its
elastic modulus and strength, is modeled by linearly interpo-
lating the experimental data appearing in Table 2 with time.
Because Panels LT1 and LT4 to LT9 were cast using the same
concrete mixture design, cured under similar conditions and
loaded at the same age, their material properties, including
strength, elastic modulus, and shrinkage are considered
the same and the average values of the experimental data
reported in Fig. 5 and Table 2 are used in the model. The
creep properties of concrete, however, were not averaged in
Fig. 6—Creep coefficients measured in test.

Table 3—Comparison of test results to predicted results by theoretical model


Test results Model predictions
Concrete age upon Test duration, Residual failure load Critical time for creep Short-term failure
Panel No. loading (t0), days days Creep buckling after creep PR (kN) buckling*, days load Pu†, kN PR/Pu
LT1 22 13 √ 0 18 638 0
LT2 146 57 — 373 N/A 436 86%
LT3 99 42 — 794 N/A 820 97%
LT4 22 104 — 466 N/A 499 93%
LT5 22 69 √ 0 157 764 0
LT6 22 112 — 665 N/A 747 89%
LT7 22 112 — 1377 N/A 1526 90%
LT8 22 82 √ 0 56 1301 0
LT9 22 8 √ 0 11 1075 0
*
Prediction is based on long-term analysis presented in Huang and Hamed.16

Prediction is based on short-term analysis presented in Huang et al.7
Note: 1 kN = 0.225 kip.

46 ACI Structural Journal/January 2018


Table 4–Moduli of springs for Maxwell model of each panel
Panel No. E1, MPa E2, MPa E3, MPa E4, MPa E5, MPa E6, MPa
LT1 9313 4573 3982 3097 2889 15,200
LT2 3494 3661 2560 4849 5493 20,058
LT3 3494 3661 2560 4849 5493 20,058
LT4 8247 4428 3890 3043 2845 16,662
LT5 8247 4428 3890 3043 2845 16,662
LT6 4588 1593 4463 5157 1820 21,882
LT7 4588 1593 4463 5157 1820 21,882
LT8 8224 6047 4398 3717 3628 12,866
LT9 8224 6047 4398 3717 3628 12,866

Note: 1 MPa = 0.145 ksi.

the model due to the notable differences obtained from the


control tests of different batches.
A generalized Maxwell model20 was used in Hamed and
Huang16 to model the concrete following the expansion
of the relaxation function into a Dirichlet series. It was
assumed that the relaxation function is approximately equal
to the inverse of the compliance function of concrete—that
is, Ec/(1 + φc), where Ec is the elastic modulus of concrete
and φc is the creep coefficient. Thus, the experimental relax-
ation function can be obtained by substituting the measured
elastic modulus and creep coefficient into the formula at
different times. By using the least-squares method to fit
the experimental relaxation function, the spring moduli in Fig. 7—Representative buckling failure mode.
the Maxwell chain are determined. Five Maxwell units are panels exhibited distinct time-dependent behaviors under
used to model the viscoelastic behavior of concrete with τμ the influence of creep and shrinkage. Panel LT6 exhibited
= 5μ–1 (days), where τμ is the relaxation time of the μ-th unit. a gradually decreasing rate of increase of deformations over
The spring constants that are obtained by the least-squares time with a stable time-dependent response. Panels LT1 and
methods, which were used for the analysis of the panels, are LT5, on the other hand, experienced three stages of increase
given in Table 4. in the deflections and failed by creep buckling at 13 days and
Cracking occurred in all panels during initial loading 69 days since initial loading, respectively. At the initial stage
or within a half-day after the imposed instantaneous load after the sustained load was applied to Panels LT1 and LT5,
reached the desired sustained load level. The cracks first the deflections increased at a decreasing rate. Following this,
started in the middle span and then propagated toward the a transitional stage occurred where the rate of increase in
ends. The cracked regions were generally concentrated the deflections remained nearly constant. In the third stage,
within the middle one-third span of the panels at the time of the deflections of the panels increased at an accelerating rate
failure. Panels LT1, LT5, LT8, and LT9 failed by creep buck- (so-called tertiary creep response), which ultimately led to
ling under the sustained loads at different times, as shown buckling failure of the panels.
in Table 3. The other panels exhibited stable responses It is clear that a higher sustained load level leads to a more
with time and were then further loaded to failure. The residual rapid increase of deflection as well as a higher tendency
failure loads of Panels LT2 to LT4 and Panels LT6 to LT7 at the of creep buckling for slender HSC panels. It can be seen
end of the sustained loading tests are also given in Table 3. As from Fig. 8 that the theoretical results for all three panels
observed from the strain gauges readings, all panel speci- correlate with the test results reasonably well. The esti-
mens collapsed in a buckling failure mode before strain mated critical time for creep buckling failure is 18 days and
softening of the concrete. A typical buckling failure mode is 157 days for Panels LT1 and LT5, respectively. The discrep-
shown in Fig. 7. ancy between the experimental and predicted critical times
for Panel LT5 is believed to be due to the large sensitivity of
Deflection versus time the panel response to uncertainties in the actual load level.
The variation of the measured and predicted central To strengthen this argument, Fig. 9 shows that increasing the
deflections with time for all Panels except LT4 are presented load level by 5% only, which is within the typical tolerance
in Fig. 8. Panels LT1, LT5, and LT6 were loaded at the same of an experimental test, can dramatically change the critical
age (22 days) but under different load levels, which were time that leads to creep buckling, and yields a better agree-
93%, 72%, and 66% of the corresponding short-term failure ment between the theoretical predictions and the test results.
loads, respectively. It can be seen in Fig. 8 that the three

ACI Structural Journal/January 2018 47


Fig. 8—Time variation of out-of-plane deflections of center points of Panels LT1 to LT3 and LT5 to LT9. (Note: 1 mm = 0.0394 in.)
age of concrete, the early-loaded Panel LT5 showed a larger
increase with time in the deflection and exhibited a sudden
buckling failure after 69 days, whereas Panel LT3 exhibited
a stable response. Hence, to reduce the risk of creep buckling
failures in slender HSC panels, it is desirable to avoid an
early-age loading.
Panels LT7 and LT8 were both tested at 22 days with
different sustained load levels that equal 53% and 71% of
their corresponding short-term load capacities, respectively.
The designed eccentricities were the same and equal to h/12.
The deflection of Panel LT7 reached a limit, which revealed
a stable response, whereas Panel LT8 failed at 82 days by
creep buckling. The model results correlated well with the
test results. One other reason that could have contributed to
Fig. 9—Illustration of sensitivity of panel behavior to load the differences in the behavior is that Panel LT8 was made
level for Panel LT5. (Note: 1 mm = 0.0394 in.) from concrete Batch 5, which exhibits more creep than
Panel LT2 was tested at the age of 146 days with a designed Batch 4, from which Panel LT7 was made (refer to Fig. 6).
eccentricity of h/3 and a sustained load level of 73% of its Panel LT9 had a thickness of 130 mm (5.12 in.) (slender-
short-term load capacity. Panel LT3 was tested at the age ness ratio equal to 72) in contrast to the other panels whose
of 99 days under a sustained load of 75% of its short-term thickness was 100 mm (3.94 in.) (slenderness ratio equal to
load capacity with an eccentricity of h/6. It can be seen in 94). It was loaded to 95% of the short-term failure load with
Fig. 8 that both panels displayed increased deflections over an eccentricity of h/6 at the age of 22 days. The panel expe-
time at decreasing rates. The deflection-versus-time curves rienced three stages of increase in the deflection and failed
asymptotically approached constant values, showing a stable quickly by creep buckling at 8 days after initial loading. It
state. The ratios of the deflections at the end of the test over can be seen in Fig. 8 that the theoretical model describes the
the instantaneous deflections for Panels LT2 and LT3 were time-dependent behavior very well, with a predicted critical
1.66 and 1.35, respectively, whereas the creep coefficients time of 11 days.
of the corresponding concrete cylinders were 0.41 and 0.39, The effect of the slenderness ratio can be clarified by
respectively. Thus, the deflections in the panels underwent comparing the results of Panels LT1 and LT9, as they were
a more rapid increase than the creep coefficients, owing to both loaded under similar load levels and load eccentricities.
the contributions of the geometric nonlinearity (P-Δ effect) In addition, they were made from concrete Batches (1 and
and the cracking of concrete. It can be seen in Fig. 8 that the 5, respectively) that exhibited more or less the same creep
predictions by the theoretical model agree well with the test response (refer to Fig. 6). It is seen in Fig. 8 that Panel LT9,
results for Panels LT2 and LT3. which is less slender than Panel LT1, buckled more quickly.
The effect of aging can be examined by comparing the This observation is contrary to the expectation where the
results of Panel LT3 and Panel LT5, which had similar load more slender panel should buckle more rapidly. This is
levels and eccentricities but were loaded at 99 days and explained by the difference in the sustained load ratio: LT9
22 days, respectively. It can be observed in Fig. 8 that due had slightly 2% larger load level than LT1. Although it is
to the strong dependency of the creep and shrinkage on the a very small difference, as mentioned previously, the creep

48 ACI Structural Journal/January 2018


buckling response of HSC panels is very sensitive to the buckling failure for Panel LT1 is approximately 60% of the
magnitude of the applied sustained loading, especially at instantaneous failure load after 3087 days since loading at the
high load levels. Therefore, an additional case is analyzed age of 22 days. Thus, the buckling capacity of HSC panels
where the load level in Panel LT9 is reduced by 2% to 93% can be reduced by 40% because of the combined effects of
of its short-term load-carrying capacity. The response is creep, shrinkage, and cracking. The reductions may vary
given in Fig. 10. The behavior of LT9 under a load level for different boundary conditions, slenderness ratios, and
of 93% significantly differs from that under a load level of reinforcement ratios, which should be carefully considered
95%. The increase in the out-of-plane deflection is much in the design and analysis of slender HSC panels.
slower when the load level is 93% and the panel buckles at
292 days. Thus, if the two panels (LT1 and LT9) were loaded Strains versus time
by exactly the same load levels, then the more slender panel The measured and theoretical strains of concrete and steel
(LT1) would fail first. reinforcement at the bottom face of the midlength section for
The theoretical model, which is validated herein, predicts the panels that failed by creep buckling (Panels LT1, LT5,
that the minimum sustained load that will lead to creep LT8 and LT9) are plotted against time in Fig. 11. Similar
to the deflection, the compressive strains of the concrete
and steel also increase with time as a result of the combined
effects of creep, shrinkage, and cracking of concrete, as well
as the geometric nonlinearity. Moreover, the concrete strains
are far below the measured peak strain of 3500 µε that corre-
sponds to the compressive strength throughout the entire
creeping period. This means no strain softening of concrete
occurred until failure. The strains in the steel reinforcement
are also smaller than the yielding strain of 2000 µε during
most of the creeping period. These measurements, along
with the development of tertiary creep deformations (Fig. 8),
emphasize that buckling is the failure mode and not typical
flexural failure. In some measurements, it can be seen that
the strains suddenly increased to exceed the limit strains of
Fig. 10—Illustration of sensitivity of panel behavior to load the steel only at the moment of failure (LT9, for example).
level for Panel LT9. (Note: 1 mm = 0.0394 in.)

Fig. 11—Time variation of strains in concrete and steel reinforcement at bottom face of midlength section of Panels LT1, LT5,
LT8, and LT9.

ACI Structural Journal/January 2018 49


Fig. 12—In-plane load versus out-of-plane center deflection of Panels LT2, LT3, LT4, LT6, and LT7. (Note: 1 kN = 0.225 kip;
1 mm = 0.0394 in.)
Figure 11 shows that the predicted strains of concrete and LT6, and LT7 are approximately 10 to 14% smaller than the
steel have good correlation with the test results for Panels predicted short-term capacities, but only a 3% reduction in
LT1 and LT5. For Panels LT8 and LT9, a reasonably good the load-carrying capacity is obtained for Panel LT3. Based
agreement is achieved for concrete, but there exist some on these results, it is concluded that the long-term effects of
discrepancies for the strains in the steel. The reasons for the creep and shrinkage could lead to reductions in the residual
discrepancies are not quite clear, but they might be related to strength of HSC panels to varying degrees, which need to be
local dislocation of the steel mesh or unusual strain drifting considered in their design.
due to poor installation of the strain gauges on the steel As mentioned previously, Panel LT4 was first loaded to
reinforcement. 93% of its predicted short-term failure load. The load was
then dropped to 74% of its failure load, and then it was kept
Load versus deflection constant with time. The panel exhibited a stable behavior
As seen in Fig. 8, Panels LT2, LT3, LT4, LT6, and LT7 during the duration of the sustained loading test, and exhib-
were stable during the sustained loading tests. There- ited a small reduction of approximately 7% in the strength
fore, these panels were loaded to failure at the end of the when loaded to failure after creep.
sustained loading tests by increasing the load without
releasing the sustained load. The experimental loads versus CONCLUSIONS
the out-of-plane deflections of center points of the Panels are The outcomes of an experimental study that consisted
plotted in Fig. 12. The figure also shows the theoretical load- of testing nine slender one-way HSC wall panels under
versus-deflection curves produced by the short-term model.7 sustained loading is reported in this paper. The panels were
The theoretical results used the material properties of tested subjected to in-plane eccentric compression loads and were
specimens at the age of initial loading and excluded the simply supported along their short edges. Four panels failed
effects of creep and shrinkage. The deflections of all panels by creep buckling under sustained loads at certain times,
increased almost linearly with the loads at the initial stage while the other five panels exhibited stable response over
of loading. At the sustained load levels, the loads were kept time and were then loaded to failure to determine their
nearly constant over time and the deflections increased due residual capacities. Three parameters that can substantially
to the effects of creep and shrinkage. Figure 12 shows that affect the response of the panels are considered: 1) the
the magnitudes of the deflections after increasing the load loading age; 2) the magnitude of the in-plane load; and 3)
to failure were larger than those predicted under short- the slenderness ratio.
term loading. The ratios of the residual failure loads over While the loading age is known to influence the creep of
the predicted instantaneous failure loads are given in Table concrete, this parameter seems to have a crucial influence in
3. It can be seen that the residual capacities of Panels LT2, geometrically nonlinear buckling problems, as it may domi-

50 ACI Structural Journal/January 2018


nate the stability of HSC panels. The panel that was tested at REFERENCES
99 days under the same loading conditions as its counterpart 1. Bažant, Z. P., “Creep Stability and Buckling Strength of Concrete
Columns,” Magazine of Concrete Research, V. 20, No. 63, 1968, pp. 85-94.
panel, which was tested at 22 days, showed a stable, time- doi: 10.1680/macr.1968.20.63.85
dependent response. On the other hand, creep buckling 2. Bockhold, J., and Petryna, Y. S., “Creep Influence on Buckling Resis-
failure occurred to the panel that was loaded at 22 days. tance of Reinforced Concrete Shells,” Computers & Structures, V. 86,
No.  7-8, 2008, pp. 702-713. doi: 10.1016/j.compstruc.2007.07.004
Therefore, it is desirable to avoid loading at early ages to 3. Hamed, E.; Bradford, M. A.; Gilbert, R. I.; and Chang, Z. T., “Analytical
reduce the risk of creep buckling failures of slender HSC Model and Experimental Study of Failure Behavior of Thin-Walled Shallow
panels. Furthermore, the results reveal that the long-term Concrete Domes,” Journal of Structural Engineering, ASCE, V. 137, No. 1,
2011, pp. 88-99. doi: 10.1061/(ASCE)ST.1943-541X.0000274
behavior of HSC panels is very sensitive to the in-plane load 4. Oberlender, G. D., and Everard, N. J., “Investigation of Reinforced
level. It is found that a small difference in the load level such Concrete Walls,” ACI Journal Proceedings, V. 74, No. 6, June 1977,
as 2% can give rise to notable variation in the critical time pp. 256-263.
5. Saheb, S. M., and Desayi, P., “Ultimate Strength of RC Wall
that leads to creep buckling. Panels in One-Way In-Plane Action,” Journal of Structural Engi-
The time-dependent effects of creep and shrinkage, coupled neering, ASCE, V. 115, No. 10, 1989, pp. 2617-2630. doi: 10.1061/
with the geometric nonlinearity (P-Δ effect), have caused (ASCE)0733-9445(1989)115:10(2617)
6. Fragomeni, S., and Mendis, P. A., “Instability Analysis of Normal-
reductions in the load-carrying capacity of HSC panels to and High-Strength Reinforced Concrete Walls,” Journal of Structural
varying degrees. Some panels failed with time by creep buck- Engineering, ASCE, V. 123, No. 5, 1997, pp. 680-684. doi: 10.1061/
ling under sustained loads that were approximately 71% of (ASCE)0733-9445(1997)123:5(680)
7. Huang, Y.; Hamed, E.; Chang, Z. T.; and Foster, S. J., “Theoretical and
their short-term load-carrying capacity. Panels that exhibited Experimental Investigation of Failure Behavior of One-Way High-Strength
a stable response with time were loaded to failure after creep Concrete Wall Panels,” Journal of Structural Engineering, ASCE, V. 141,
and showed a reduction of up to 14% in their residual capacity. No. 5, 2015, 4014143. doi: 10.1061/(ASCE)ST.1943-541X.0001072
8. Green, R., and Breen, J. E., “Eccentrically Loaded Concrete Columns
Predictions based on the theoretical model of Huang and under Sustained Load,” ACI Journal Proceedings, V. 66, No. 11, Nov. 1969,
Hamed16 that was validated in this study show that sustained pp. 866-874.
loads as low as 60% of the short-term capacity can lead 9. Goyal, B. B., and Jackson, N., “Slender Concrete Columns under
Sustained Load,” Journal of the Structural Division, ASCE, V. 97, 1971,
to creep buckling failures. Such reductions in the load-car- pp. 2729-2750.
rying capacity caused by the long-term effects of creep and 10. Behan, J. E., and O’Connor, C., “Creep Buckling of Reinforced
shrinkage should be carefully considered and treated in the Concrete Columns,” Journal of the Structural Division, ASCE, V. 108,
1982, pp. 2799-2818.
design of slender HSC wall panels. 11. Mickleborough, N. C., and Gilbert, R. I., “Creep Buckling of Uniax-
ially Loaded Reinforced Concrete Columns,” Computer Analysis of Effects
AUTHOR BIOS of Creep, Shrinkage, and Temperature Changes on Concrete Structures,
Yue Huang is a Research Associate at the Center for Infrastructure Engi- SP-129, C. C. Fu and M. D. Daye, eds., American Concrete Institute, Farm-
neering and Safety, School of Civil and Environmental Engineering, the ington Hills, MI, 1991, pp. 39-55.
University of New South Wales, Kensington, NSW, Australia. He received 12. Claeson, C., and Gylltoft, K., “Slender Concrete Columns Subjected
his BEng in structural engineering from City University of Hong Kong, to Sustained and Short-Term Eccentric Loading,” ACI Structural Journal,
Hong Kong, and his PhD from the University of New South Wales. His V. 97, No. 1, Jan.-Feb. 2000, pp. 45-53.
research interests include time-dependent behavior of reinforced concrete 13. Bradford, M. A., “Shrinkage and Creep Response of Slender Rein-
structures. forced Concrete Columns under Moment Gradient: Theory and Test
Results,” Magazine of Concrete Research, V. 57, No. 4, 2005, pp. 235-246.
ACI member Ehab Hamed is a Senior Lecturer at the Center for Infra- doi: 10.1680/macr.2005.57.4.235
structure Engineering and Safety, School of Civil and Environmental Engi- 14. Tatsa, E. Z., “Load Carrying of Eccentrically Loaded Reinforced
neering, the University of New South Wales. He received his PhD in struc- Concrete Panels under Sustained Load,” ACI Structural Journal, V. 86,
tural engineering from the Technion – Israel Institute of Technology, Haifa, No. 2, Mar.-Apr. 1989, pp. 150-155.
Israel. His research interests include the time-dependent behavior of materials 15. Lee, Y.; Lee, B. Y.; Kwon, S. H.; Kim, Y. Y.; and Kim, J. K.,
and structures as well as the repair of steel and concrete structures using “Long-Term Behavior of a Reinforced Concrete Wall under Compressive
composite materials. Stress Applied to Part of the Wall’s Entire Width,” Magazine of Concrete
Research, V. 60, No. 10, 2008, pp. 759-768. doi: 10.1680/macr.2008.00090
Zhen-Tian Chang is a Senior Research Fellow at the Center for Infra- 16. Huang, Y., and Hamed, E., “Buckling of One-Way High-Strength
structure Engineering and Safety, School of Civil and Environmental Engi- Concrete Panels: Creep and Shrinkage Effects,” Journal of Engineering
neering, the University of New South Wales. He received his BEng and Mechanics, ASCE, V. 139, No. 12, 2013, pp. 1856-1867. doi: 10.1061/
MEng in structural engineering from Hunan University, Changsha, China, (ASCE)EM.1943-7889.0000629
and his PhD from the University of New South Wales. His research inter- 17. Fragomeni, S., and Mendis, P. A., “Applicability of Current
ests include steel and reinforced concrete structures and durability of infra- ACI 318 Wall Design Formula for High Strength Concrete Walls,”
structures in aggressive environments. Advances in Structural Engineering, V. 2, No. 2, 1999, pp. 103-108. doi:
10.1177/136943329900200203
ACI member Stephen J. Foster is Professor and Head of the School of 18. Standards Association of Australia, “Methods of Testing Concrete—
Civil and Environmental Engineering at the University of New South Wales, Method 13: Determination of the Drying Shrinkage of Concrete for
where he received his PhD in 1993. His research interests include the struc- Samples Prepared in the Field or in the Laboratory,” AS 1012.13, Sydney,
tural use of high-strength and ultra-high-strength concretes, fiber-reinforced Australia, 1992.
concrete, nonlinear finite element analysis of concrete structures, and reha- 19. Fields, K., and Bischoff, P. H., “Tension Stiffening and Cracking
bilitation of concrete structures. of High-Strength Reinforced Concrete Tension Members,” ACI Structural
Journal, V. 101, No. 4, July-Aug. 2004, pp. 447-456.
20. Bažant, Z. P., and Wu, S. T., “Rate-Type Creep Law of Aging
ACKNOWLEDGMENTS Concrete Based on Maxwell Chain,” Materials and Structures, V. 7,
The work reported in this paper was supported by the Australian Research No. 1, 1974, pp. 45-60.
Council (ARC) through a Discovery Project (DP120102762).

ACI Structural Journal/January 2018 51


JOIN AN
ACIChapter!
The American Concrete Institute has Chapters and Student Chapters located
throughout the world. Participation in a local chapter can be extremely
rewarding in terms of gaining greater technical knowledge and networking
with leaders in the concrete community.
Because chapters are distinct and independent legal entities, membership includes both ACI
members and non-ACI members and is made up of a diverse blend of architects, engineers,
consultants, contractors, educators, material suppliers, equipment suppliers, owners, and
students—basically anyone interested in concrete. Many active ACI members initially became
involved in ACI through their local chapter. In addition to technical programs and publications,
many chapters sponsor ACI Certification programs, ACI educational seminars, project award
recognition programs, and social events with the goal of advancing concrete knowledge.

Check out the Chapters Special Section from the October 2015 Concrete International:
www.concrete.org/publications/concreteinternational/cibackissues.aspx?m=456

Student Chapters
Join or form an ACI Student Chapter to maximize your influence, knowledge
sharing, and camaraderie! ACI has over 109 student chapters located
throughout the world, each providing opportunities for students to:
• Connect with their peers and participate in concrete-related activities such as: student
competitions, ACI Conventions, ACI Certification Programs, ACI Educational Seminars, local
chapter meetings, social events, and community service projects;
• Network with members of local chapters, many of whom have been in the industry for
decades and can help to develop professional relationships and offer career advice;
• Win recognition for their universities through the University Award; and
• Learn about the many scholarships and fellowships offered by the ACI Foundation and by
ACI’s local chapters.

ACI Student Chapters worldwide

   
www.concrete.org/chapters
ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 115-S04

Nonlinear Backbone Modeling of Concrete Columns


Retrofitted with Fiber-Reinforced Polymer or Steel Jackets
by José C. Alvarez, Sergio F. Breña, and Sanjay R. Arwade

The use of nonlinear analysis procedures in seismic retrofitting of hinges form at the ends of beams first instead of at the ends
existing concrete structures has become commonplace in practice. of columns. This causes the collapse mechanism to change
Backbone curves are often used to capture the nonlinear response from a failure involving limited lateral deformation capacity
of components in a simplified but sufficiently accurate way. Proce- to one involving a more ductile mechanism. As a local
dures to construct curves for existing components of frames
retrofitting technique, jacketing can be selectively used in
(beams, joints, and columns) have received considerable attention
cases where other components of frames that make up the
and have been modified over the years. In contrast, recommenda-
tions to construct these backbone curves for retrofitted components lateral load system have sufficient strength and deformation
are largely lacking. This paper presents recommendations that capacity. It is critical, however, to be able to calculate the
can be used to construct backbone curves of circular and rect- expected sequence of hinge formation by developing reli-
angular retrofitted columns using jacketing materials within the able nonlinear modeling techniques of retrofitted elements
context of ASCE/SEI 41-13 and ACI 369R-11. The recommenda- of potentially vulnerable reinforced frames.
tions are developed using a sectional model for force parameters This paper focuses on developing a methodology to model
and a statistical study of load-deformation results from a database reinforced concrete columns that are retrofitted using steel
of retrofitted columns for drift parameters when sufficient experi- or FRP jackets applied locally in the region where inelastic
mental data are available. Key points in the backbone response of hinges are anticipated to form. The methodology focuses
jacketed columns are summarized in tabular form consistent with
on a recommended procedure to construct the nonlinear
ASCE/SEI 41-13 to facilitate possible adoption in future updates of
force-deformation backbone response of jacketed columns.
the standard. A probabilistic model is introduced to allow selection
of drift values that correspond to selected exceedance probabilities. The procedures used within this methodology result in
backbone curves that compare favorably with the backbone
Keywords: backbone curves; column retrofitting; jacketed columns; response measured in available tests. Recommendations to
nonlinear modeling parameters. construct backbone curves are provided in a similar format
to the current approach for existing elements contained in
INTRODUCTION ASCE/SEI 41-13 and ACI 369R-11 by defining key points
The potential for failure of nonductile columns with in the nonlinear backbone response of reinforced concrete
details that do not conform to modern building codes has frame elements. The recommendations were developed in
been recognized for many years. With local retrofitting such a way that they may provide the basis for future updates of
as external jacketing, the behavior of nonductile columns these two documents.
can be greatly improved, providing benefits to the global The backbone curves are defined in this paper by six
performance of building frames and reducing the potential parameters: the force at yield, the peak force, the residual
for collapse. Jacketing materials are selected such that they capacity, the drift at yield, the drift at 80% of peak force,
can be easily applied to frame elements without disruption and the drift at which the force equals the residual capacity.
of building operations. The most common types of external A sectional model, validated against a database of experi-
jackets are concrete, steel, and fiber-reinforced polymer mental results, is used to calculate the force at yield and the
(FRP) materials. peak force. Validated sectional models that can predict drift
In any retrofitting project, it is paramount to be able to of jacketed columns are not currently available and, there-
accurately assess the performance of the retrofitted structure fore, a statistical model for the drift at 80% of peak force
to ensure adequacy in its performance. ASCE/SEI 41-13, is developed and calibrated to a database of experimental
“Standard on Seismic Evaluation and Retrofit of Existing results assembled from the open literature for cases when
Buildings,” provides detailed guidance for the evaluation of the axial force ratio is low. Unfortunately, insufficient data
structures in their existing condition. In contrast, very little are available in the literature to develop a statistical model
information is provided for engineers to verify the response for the drift at 80% of peak load when the axial force ratio
of the structure in its retrofitted condition. Guidance on is high or for drift corresponding to the residual capacity of
procedures to estimate strength and deformation capacity of
retrofitted components is needed to adequately evaluate if ACI Structural Journal, V. 115, No. 1, January 2018.
the structure satisfies the intended performance objectives MS No. S-2016-121.R3, doi: 10.14359/51700779, received March 10, 2017, and
reviewed under Institute publication policies. Copyright © 2018, American Concrete
for the expected demands. Institute. All rights reserved, including the making of copies unless permission is
Jacketing deficient columns can change the sequence obtained from the copyright proprietors. Pertinent discussion including author’s
closure, if any, will be published ten months from this journal’s date if the discussion
of inelastic hinge formation of an existing frame such that is received within four months of the paper’s print publication.

ACI Structural Journal/January 2018 53


reduction in lateral strength from peak. Parameter b is taken
as the plastic drift that corresponds to the point where the
lateral strength has degraded to the residual strength, gener-
ally assumed equal to 0.2Vpeak. These two definitions are
followed to construct simplified backbone curves using a
database of existing tests of jacketed columns as discussed
in the following section.

DATABASE OF JACKETED COLUMNS


An experimental database of available laboratory exper-
iments of circular and rectangular jacketed columns tested
under cyclic loading was assembled for two main purposes.
First, lateral load data were used to evaluate the accuracy
of proposed sectional models described later in this paper
Fig. 1—Simplified backbone curve. to compute yield and peak strength of jacketed columns.
Second, the data were also used to study the statistical
jacketed columns. Those parameters are therefore conser- distribution of drifts at key points of the force-deformation
vatively estimated as for unjacketed columns containing response of jacketed columns. The information obtained
details similar to those in modern codes. from the database, therefore, was used to evaluate the accu-
racy of the methods proposed to construct the backbone
RESEARCH SIGNIFICANCE response of jacketed columns.
ASCE/SEI 41-13 and ACI 369R-11 contain requirements The database consists of a total of 116 columns: 84 and
to model existing reinforced concrete building frames for 32 columns jacketed using fiber-reinforced polymer (FRP)
retrofitting purposes. These documents focus on modeling or steel materials, respectively. Several challenges were
of existing components but do not provide details on ways encountered when assembling the database, given the vast
to treat retrofitted components. This paper is intended to fill variety of details in the columns tested. The specimens were
this gap in information by developing a procedure that can designed and constructed to capture common deficient prop-
be used to construct nonlinear backbone curves for jacketed erties typically found in columns of older existing build-
columns, a technique that is frequently considered in retro- ings such as low shear strength, short lap splices within
fitting projects. the plastic hinge zone, and/or low volume of transverse
reinforcement. Column jacket retrofitting using different
GENERAL BACKBONE FORCE-DEFORMATION materials and configurations was used to overcome these
RELATIONSHIP existing deficiencies that result in poor column performance.
Both ASCE/SEI 41-13 and ACI 369R-11 use a simpli- The laboratory specimens include columns with rectan-
fied backbone model to describe the nonlinear lateral load- gular or circular cross sections, various levels of applied
drift response of existing reinforced concrete elements axial force (typically axial load levels below the balanced
(Fig. 1). In this model, the nonlinear response is primarily point in the column interaction diagram), and two different
characterized through the definition of two parameters (a test setup configurations that generate single- or double-
and b) that are used to represent the plastic drift capacity curvature bending of the columns during lateral loading.
of existing columns for given transverse reinforcement Relevant properties of jacketed columns in the database are
contents and axial load level. These documents tabulate given in Table 1 for circular columns and Table 2 for rect-
values of nonlinear drift parameters for columns and other angular columns, respectively. Many (not all) of the tests on
reinforced concrete components based on experience and retrofitted columns were compared with tests of an unretro-
the observed response during laboratory testing of compo- fitted reference column, but only the retrofitted specimens
nents. One of the drawbacks in the currently used simplified are provided in the tables. It can be seen that the highest
backbone model in comparison with response of compo- number of specimens correspond to FRP-jacketed columns
nents measured during tests is that the strength drop after with rectangular cross sections and the fewest number of
reaching peak strength is not as sudden, as indicated by the specimens obtained correspond to steel-jacketed circular
solid line in the figure, but it is rather gradual in most cases columns. The range of values for key column properties in
(dashed line). The simplified backbone response curves for the database can also be found within these tables (material
jacketed columns proposed later in this paper will follow the strength, reinforcement content, dimensions).
general shape as indicated by the dashed line to better reflect The original (unretrofitted) columns contained one of
observed response. the primary three deficiencies that result in nonductile
Because of the widespread use of simplified back- lateral-load behavior of columns: short lap splices in the
bone curves in nonlinear analysis of existing structures, plastic hinge region (LS); insufficient amounts of trans-
the approach followed in this paper is to characterize the verse reinforcement for concrete confinement (C); or low
nonlinear lateral deformation response of jacketed columns shear strength (S). The measured lateral load-deformation
through parameters a and b. Parameter a can be defined as response of the column specimens was reported in the orig-
the magnitude of plastic drift that corresponds to a 20% inal reference through either hysteresis curves or backbone

54 ACI Structural Journal/January 2018


Table 1—Properties of jacketed columns in database: circular sections
Height, Deficient Δy/H, Δp/H, Δmax/H, Par.
Specimen Reference Dc, in. in. fc′, psi P/Ag fc′ ρ, % ρv, % property* Vy, kip Vpeak/Vy % % % a
FRP-jacketed circular columns
CFRP-05 9.5 37.5 3467 0.05 2.54 0.333 LS 9.4 1.4 1.9 6.0 9.2 6.1
KFRP-05 Breña and 9.5 37.5 3467 0.05 2.54 0.333 LS 9.6 1.4 1.4 6.3 9.5 6.8
Schlick
CFRP-15 (2007) 9.5 37.5 3467 0.15 2.54 0.333 LS 9.1 1.8 1.3 6.0 9.2 7.9
KFRP-15 9.5 37.5 3467 0.15 2.54 0.333 LS 9.5 1.4 1.1 6.2 10.0 8.8
CF-R1 24.0 135.0 5223 0.06 1.95 0.103 LS 25.4 1.4 0.4 5.3 6.4 6.0
CF-R2 24.0 135.0 5353 0.06 1.95 0.103 LS 28.3 1.4 0.4 4.4 5.9 5.5
CF-R3 24.0 135.0 4758 0.07 1.95 0.103 LS 31.2 1.4 0.5 3.8 5.6 5.1
CF-R4 24.0 135.0 5469 0.06 1.95 0.103 LS 30.7 1.4 0.4 3.9 5.9 5.5
CF-R5 24.0 135.0 5759 0.06 1.95 0.103 LS 31.1 1.3 0.5 3.9 5.9 5.4
Haroun and
CF-R6 Elsanadedy 24.0 135.0 4802 0.07 1.95 0.103 LS 30.3 1.4 0.4 4.8 7.0 6.6
(2005a,b)
CS-R1 24.0 96.0 5919 0.05 1.95 0.103 S 80.2 1.4 0.2 3.1 3.6 3.4
CS-R2 24.0 96.0 5687 0.05 1.95 0.103 S 83.2 1.4 0.2 3.0 3.7 3.5
CS-R3 24.0 96.0 4961 0.05 1.95 0.103 S 116.1 1.4 0.5 3.8 4.3 3.8
CS-R4 24.0 96.0 5455 0.06 1.95 0.103 S 112.1 1.4 0.5 2.1 3.1 2.5
CS-P1 24.0 96.0 5179 0.06 1.95 0.103 S 119.8 1.4 0.3 3.9 4.1 3.7
Priestley et
Lap Splice R 24.0 144.0 4998 0.18 2.53 0.103 LS 42.7 1.4 0.1 2.4 3.8 3.5
al. (1994a,b)
C2-RT4 24.0 104.0 6501 0.05 1.94 0.103 LS 53.2 1.3 0.4 1.4 5.1 3.2
Xiao and
C3-RT5 24.0 104.0 6501 0.05 1.94 0.103 LS 52.7 1.4 0.4 2.5 5.0 3.7
Ma (1997)
C4-RP4 24.0 104.0 6501 0.05 1.94 0.103 LS 36.5 1.4 0.6 2.0 5.4 2.7
CAF1-2N 14.0 57.9 3611 0.05 1.71 0.275 LS 17.0 1.3 1.0 1.2 4.9 1.1
CAF1-5N Ghosh and 14.0 57.9 3640 0.27 1.71 0.275 LS 12.2 1.4 0.8 4.3 12.3 5.1
Sheikh
CBF1-6N (2007) 14.0 57.9 3843 0.05 1.71 1.032 LS 13.6 1.3 1.0 5.5 8.7 6.2
ST-4NT 14.0 57.9 6497 0.27 1.71 1.032 LS 25.7 1.3 1.8 3.9 9.0 7.2
Steel-jacketed circular columns
2 24.0 144.0 5601 0.16 2.53 0.174 C 38.8 1.4 0.8 2.5 2.5 1.7
4 24.0 144.0 5521 0.16 2.53 0.174 C 48.0 1.4 0.4 6.0 †
5.6
Chai et al.
5 24.0 144.0 5095 0.17 2.53 0.174 C 36.8 1.3 0.4 1.1 6.0 2.9
(1991)
6 24.0 144.0 5426 0.16 2.53 0.174 C 49.2 1.4 0.5 4.6 6.1 5.4
1-R 24.0 144.0 5541 0.16 2.53 0.174 C 38.0 1.4 0.9 2.7 5.1 3.9
SC1 Li et al. 29.9 128.0 3699 0.11 1.32 0.072 C 54.9 1.4 0.4 3.8 5.8 4.5
SC2 (2005) 29.9 128.0 3699 0.11 1.15 0.067 C 51.3 1.4 0.3 3.8 5.8 5.5
C2R 24.0 96.0 4931 0.06 2.53 0.082 S 115.4 1.4 0.3 4.4 ‡ 4.1
C4R Priestley et 24.0 96.0 5101 0.17 2.53 0.082 S 150.4 1.4 0.3 4.1 ‡ 3.8
C6R al. (1994a,b) 24.0 96.0 5801 0.05 2.53 0.082 S 160.9 1.4 0.4 5.5 ‡ 5.1
C8R 24.0 72.0 4521 0.06 2.53 0.082 S 193.0 1.4 0.3 5.2 ‡ 4.9
*
S is shear-deficient; C is inadequate confinement; LS is short lap splice.

Test was stopped at peak load.

Test stopped at maximum displacement capacity of actuator.
Notes: 1 in. = 25.4 mm, 1 kip = 4.45 kN, 1000 psi = 6.89 MPa.

curves as a measure of column performance. When only with those available in retrofitting guides and standards such
hysteresis curves were published, the backbone curve was as ACI 369R-11 and ASCE/SEI 41-13. The curves were
constructed using a series of lines passing through selected constructed by connecting the shear force and drift coor-
points in the load-deformation response curve. These back- dinates obtained during the tests corresponding to column
bone curves were subsequently simplified to be consistent yielding, lateral strength corresponding to a 20% reduction

ACI Structural Journal/January 2018 55


Table 2—Properties of jacketed columns in database: rectangular columns
bc, Deficient Δy/H, Δp/H, Δmax/H,
Specimen Reference in. hc, in. H, in. fc′, psi P/Agfc′ ρ, % ρv, % property* Vy, kip Vpeak/Vy % % % Par. a
FRP-jacketed rectangular columns
S2 15.8 7.9 70.8 1450 0.15 2.84 0.448 C 8.6 1.2 0.2 3.0 7.2 4.0
S3 Ozcan et al. 15.8 7.9 70.8 1523 0.15 2.84 0.448 C 9.4 1.3 0.3 2.0 4.8 2.7
S4 (2010) 15.8 7.9 70.8 1305 0.15 2.84 0.448 C 7.7 1.4 0.2 2.4 4.6 2.9
S5 15.8 7.9 70.8 2249 0.15 2.84 0.448 C 14.1 1.2 0.3 1.0 6.0 2.5
Confinement R Seible et al. 28.7 19.3 144.0 4998 0.14 4.65 0.128 C 107.2 1.4 0.4 2.7 3.3 2.8
Shear R (1997) 24.0 16.0 96.0 4998 0.06 2.52 0.154 S 97.9 1.2 0.1 1.5 2.3 2.1
C3 7.9 7.9 52.0 6775 0.24 2.00 0.000 C 10.8 1.3 1.0 2.1 7.2 6.2
C4 Wu et al. 7.9 7.9 52.0 6804 0.23 2.00 0.000 C 11.8 1.3 1.1 2.4 8.0 6.9
C5 (2008) 7.9 7.9 52.0 5281 0.30 2.00 0.000 C 11.1 1.3 1.0 4.0 7.2 6.2
C6 7.9 7.9 52.0 5368 0.30 2.00 0.000 C 11.0 1.3 1.0 2.0 7.3 6.3
ASG-2NSS 12.0 12.0 58.0 6165 0.15 2.44 0.316 C 24.6 1.4 0.7 4.6 23.7 2.6
ASG-3NSS 12.0 12.0 58.0 6195 0.15 2.44 0.316 C 24.9 1.4 0.2 6.0 12.0 3.5
ASG-4NSS 12.0 12.0 58.0 6282 0.15 2.44 0.316 C 24.2 1.4 0.3 2.4 11.4 2.0
Memon
ASG-5NSS and Sheikh 12.0 12.0 58.0 6340 0.15 2.44 0.316 C 24.4 1.4 0.6 4.3 11.2 2.0
(2005)
ASG-6NSS 12.0 12.0 58.0 6412 0.15 2.44 0.316 C 29.8 1.4 0.9 11.8 20.7 5.4
ASGR-7NSS 12.0 12.0 58.0 6412 0.15 2.44 0.316 C 24.2 1.4 0.5 5.8 14.9 2.6
ASGR-8NSS 12.0 12.0 58.0 6412 0.15 2.44 0.316 C 25.8 1.4 1.0 4.7 12.2 3.4
ASC-2NS 12.0 12.0 58.0 5295 0.15 2.44 0.321 C 30.4 1.3 0.3 0.7 2.3 1.3
ASC-3NS 12.0 12.0 58.0 5353 0.15 2.44 0.321 C 31.8 1.3 0.2 0.6 2.2 2.0
ASC-4NS 12.0 12.0 58.0 5353 0.15 2.44 0.321 C 26.7 1.3 0.1 0.3 2.2 0.7
Iacobucci
ASC-5NS et al. 12.0 12.0 58.0 5368 0.15 2.44 0.321 C 33.7 1.3 0.4 2.2 4.5 2.3
(2003)
ASC-6NS 12.0 12.0 58.0 5368 0.15 2.44 0.321 C 31.0 1.3 0.2 0.8 3.8 1.7
ASCR-7NS 12.0 12.0 58.0 5368 0.15 2.44 0.321 C 29.3 1.3 0.4 0.7 5.2 2.5
ASCR-8NS 12.0 12.0 58.0 6136 0.15 2.44 0.321 C 24.5 1.3 0.6 2.0 3.0 2.3
F2 18.0 18.0 70.1 3597 0.22 1.48 0.180 C 35.0 1.4 1.1 7.7 13.4 6.2
Harries et
L1 18.0 18.0 70.1 4162 0.22 1.48 0.180 LS 36.2 1.4 1.0 5.6 11.4 6.0
al. (2006)
L2 18.0 18.0 70.1 4162 0.22 1.48 0.180 LS 36.4 1.4 1.2 5.8 8.6 5.4
RF-R1 24.0 24.0 135.0 5135 0.06 2.14 0.103 LS 55.9 1.3 0.7 2.4 2.8 2.1
RF-R2 24.0 24.0 135.0 6078 0.05 2.14 0.103 LS 56.2 1.3 0.7 1.1 4.5 3.8
RF-R3 24.0 24.0 135.0 6122 0.05 2.14 0.103 LS 60.3 1.4 0.7 2.0 3.6 2.8
RF-R4 24.0 24.0 135.0 6122 0.05 2.14 0.103 LS 59.2 1.3 0.8 1.7 3.5 2.7
RS-R1 Haroun and 24.0 18.0 96.0 5527 0.06 2.04 0.137 S 107.0 1.3 0.2 1.7 4.7 3.5
Elsanadedy
RS-R2 (2005a,b) 24.0 18.0 96.0 5701 0.06 2.04 0.137 S 104.6 1.3 0.3 1.1 4.7 3.5
RS-R3 24.0 18.0 96.0 6383 0.06 2.04 0.137 S 105.1 1.3 0.2 1.5 4.4 4.1
RS-R4 24.0 18.0 96.0 6383 0.06 2.04 0.137 S 95.8 1.4 0.1 2.1 4.6 4.5
RS-R5 24.0 18.0 96.0 6383 0.06 2.04 0.137 S 97.2 1.4 0.2 2.5 4.2 4.0
RS-R6 24.0 18.0 96.0 6180 0.06 2.04 0.137 S 107.0 1.3 0.2 1.8 4.9 4.7
C1FP1 11.8 5.9 39.4 3061 0.23 1.72 0.354 LS 10.9 1.4 0.9 3.0 6.0 3.4
C1FP2 11.8 5.9 39.4 3148 0.22 1.72 0.354 LS 5.5 3.9 0.6 3.0 5.1 3.3
C1F1 11.8 5.9 39.4 3177 0.22 1.72 0.354 LS 12.5 1.3 0.9 2.0 5.0 2.5
C1F2 Harajli 11.8 5.9 39.4 3163 0.22 1.72 0.354 LS 11.2 1.4 1.0 3.0 5.1 3.4
and Rteil
C2FP1 (2004) 11.8 5.9 39.4 3061 0.27 3.56 0.354 LS 15.2 1.4 0.8 3.0 5.1 3.0
C2FP2 11.8 5.9 39.4 3148 0.26 3.56 0.354 LS 15.1 1.4 0.7 3.0 5.1 3.2
C2F1 11.8 5.9 39.4 3177 0.26 3.56 0.354 LS 15.6 1.4 0.8 3.1 5.0 3.4
C2F2 11.8 5.9 39.4 3163 0.26 3.56 0.354 LS 15.3 1.4 0.8 3.0 5.1 3.4

56 ACI Structural Journal/January 2018


Table 2(cont.)—Properties of jacketed columns in database: rectangular columns
bc, Deficient Δy/H, Δp/H, Δmax/H,
Specimen Reference in. hc, in. H, in. fc′, psi P/Agfc′ ρ, % ρv, % property* Vy, kip Vpeak/Vy % % % Par. a
SAF1-10N 12.0 12.0 57.9 3887 0.33 2.44 0.321 LS 19.9 1.3 0.8 2.5 7.8 3.2
SBF1-11N Ghosh and 12.0 12.0 57.9 3916 0.05 2.44 1.205 LS 14.3 1.3 0.8 2.3 3.8 2.1
Sheikh
SBRF1-12N (2007) 12.0 12.0 57.9 3945 0.05 2.44 1.205 LS 10.2 1.4 1.9 3.9 8.1 3.5
ASC-2NS 12.0 12.0 57.9 5294 0.33 2.44 1.205 LS 24.3 1.3 0.8 1.7 7.5 2.8
C14FP1 15.7 7.9 59.1 5656 0.00 1.29 0.735 LS 15.4 1.3 0.6 2.1 6.4 2.8
C14FP2 15.7 7.9 59.1 5656 0.00 1.29 0.735 LS 16.3 1.3 0.5 3.2 6.4 3.8
C16FP1 Harajli and 15.7 7.9 59.1 7107 0.00 2.00 0.735 LS 20.0 1.3 0.9 2.1 6.5 2.3
Dagher
C16FP2 (2008) 15.7 7.9 59.1 7107 0.00 2.00 0.735 LS 19.3 1.3 0.9 2.2 6.5 3.8
C20FP1 15.7 7.9 59.1 4641 0.00 2.13 0.735 LS 22.8 1.3 1.0 2.1 6.5 2.2
C20FP2 15.7 7.9 59.1 4641 0.00 2.13 0.735 LS 25.5 1.3 1.0 3.2 6.4 2.9
SC2 12.0 12.0 36.0 5658 0.14 6.11 0.904 S 76.6 1.3 0.1 1.8 5.7 5.6
SC1R 12.0 12.0 36.0 4932 0.16 6.11 0.904 S 72.8 1.2 0.3 1.8 5.8 3.8
SC2R Galal et al. 12.0 12.0 36.0 4932 0.16 6.11 0.904 S 76.4 1.3 0.3 1.8 4.2 2.4
SC1U (2005) 12.0 12.0 36.0 6238 0.12 6.11 0.904 S 45.4 2.2 0.1 0.9 5.6 3.2
SC3 12.0 12.0 36.0 5658 0.14 6.11 0.904 S 78.4 1.3 0.2 1.8 5.7 4.0
SC3R 12.0 12.0 36.0 4932 0.16 6.11 0.904 S 54.6 1.3 0.3 0.6 3.6 1.3
Steel-jacketed rectangular columns
C-66-R Alcocer and 19.7 19.7 78.7 4047 0.15 2.44 0.142 C 42.0 1.4 1.0 2.5 † 1.5
Durán-
C-66-S Hernández 19.7 19.7 78.7 4047 0.15 2.44 0.142 C 64.7 1.4 1.0 2.7 † 1.7
(2002)
RC-2R 10.0 10.0 40.0 8269 0.30 2.48 0.220 C 45.8 1.4 0.4 2.1 6.0 3.1
RC-3R Xiao and 10.0 10.0 40.0 8269 0.30 2.48 0.220 C 51.4 1.4 0.5 3.0 8.0 7.5
RC-4R Wu (2003) 10.0 10.0 40.0 8269 0.30 2.48 0.220 C 50.9 1.4 0.4 3.1 8.0 6.8
RC-5R 10.0 10.0 40.0 8704 0.30 2.48 0.220 C 52.7 1.4 0.4 3.1 8.5 8.1
FC9 18.0 36.0 144.0 2906 0.00 1.95 0.095 LS 38.4 1.4 0.5 2.7 4.9 3.2
FC11 Aboutaha et 18.0 36.0 144.0 2851 0.00 1.95 0.095 LS 45.4 1.4 0.8 2.4 5.5 2.5
FC12 al. (1996) 18.0 36.0 144.0 3266 0.00 1.95 0.095 LS 43.1 1.4 0.6 2.7 5.5 3.7
FC17 18.0 18.0 144.0 2636 0.00 1.95 0.076 LS 45.6 1.4 0.2 2.4 5.4 5.2
FC6 18.0 36.0 144.0 2851 0.00 1.95 0.095 LS 34.9 1.4 1.0 2.5 4.6 2.2
FC7 Aboutaha et 18.0 36.0 144.0 2981 0.00 1.95 0.153 LS 52.0 1.4 1.5 3.9 † 2.4
FC10 al. (1999b) 18.0 36.0 144.0 2596 0.00 1.95 0.095 LS 37.2 1.4 0.8 2.4 3.4 2.6
FC13 18.0 36.0 144.0 3266 0.00 1.95 0.095 LS 50.0 1.4 0.4 3.6 5.0 4.6
SC6 18.0 36.0 48.0 2256 0.00 1.95 0.095 S 111.1 1.4 1.0 3.3 5.1 3.2
SC7 Aboutaha et 18.0 36.0 48.0 2941 0.00 1.95 0.095 S 101.2 1.4 0.5 4.2 6.3 5.8
SC8 al. (1999a) 18.0 36.0 48.0 2786 0.00 1.95 0.095 S 109.7 1.4 0.7 3.9 7.0 5.8
SC10 36.0 18.0 48.0 2391 0.00 1.95 0.191 S 205.2 1.4 0.6 4.0 5.3 4.7
R2R 16.0 24.0 96.0 5601 0.05 2.52 0.163 S 104.4 1.4 0.3 3.6 ‡
Priestley
R4R et al. 16.0 24.0 96.0 5201 0.06 2.52 0.082 S 154.7 1.4 0.3 3.8 ‡
(1994a,b)
R6R 16.0 24.0 72.0 4801 0.06 2.52 0.082 S 205.6 1.4 0.4 3.7 ‡
*
S is shear deficient; C is inadequate confinement; LS is short lap splice.

Test stopped when capacity of actuator was reached.

Test stopped at maximum displacement capacity of actuator.
Notes: 1 in. = 25.4 mm; 1 kip = 4.45 kN; 1000 psi = 6.89 MPa.

from peak strength, maximum drift, and, if available, loss after loss of axial capacity and maximum drift were assumed
of axial load-carrying capacity. It should be noted that the based on experience with unjacketed columns because of
results reported in the literature were used directly without lack of experimental data reporting these values.
modification to include P-∆ effects. The residual strength

ACI Structural Journal/January 2018 57


Fig. 2—Simplified backbone curve obtained from measured hysteretic response: (a) hysteresis curve with superimposed back-
bone; and (b) simplified nonlinear backbone curve and nonlinear parameters.
The procedure used to consistently define the key points eting the region of anticipated inelastic action, but the effect
in the force-deformation response of jacketed columns is of this local retrofitting technique on the global response
illustrated schematically in Fig. 2. Shear at yield and defor- of a frame is not well understood. This section focuses on
mation at yield (Vy, Δy) were determined as the coordinates developing an approach to construct the nonlinear backbone
of the point of intersection of two lines drawn on the hyster- response of jacketed columns for use in nonlinear analyses
esis curve. One line was parallel to the initial slope of the of retrofitted frames.
measured force-deformation response of the specimens and The response of jacketed columns has been studied exten-
the second line was drawn horizontally at a shear force equal sively in laboratory tests conducted over the last 30 years.
to either 0.8Vpeak or 0.7Vpeak, depending on the shape of the These studies were initiated after dramatic failures of bridges
hysteresis curve. A shear force of 0.8Vpeak was chosen if the containing columns with deficient details occurred during
drift at peak was equal or less than 2% followed by rapid earthquakes in the two decades that followed the 1970s
strength loss at increasing drift amplitudes. Otherwise, the (for example, 1971 San Fernando Earthquake; and Hanshin
shear force at yield was assumed equal to 0.7Vpeak (Alvarez Expressway in 1994 Kobe Earthquake). Although failures of
and Breña 2014). The peak shear force Vpeak was simply taken columns in frames have not been pervasive, the reinforcing
as the highest applied lateral load during testing. Almost details that these elements contain are known to result in
all the columns in the database were not tested beyond a unsatisfactory performance under large earthquakes. The
strength degradation corresponding to a 20% drop from potentially high consequence caused by failures of defi-
Vpeak. Therefore, the residual strength of the columns was cient columns in framed buildings can be mitigated using
assumed equal to 0.2Vpeak. The force values thus determined local retrofitting—for example, by using jacketing in plastic
were used to evaluate whether the simple sectional models hinge zones near the end of columns. Development of local
described later in this paper can be used to estimate yield and retrofitting techniques by jacketing of columns was accom-
peak force of retrofitted columns. Drifts were studied statis- plished through laboratory testing, but modeling guidelines
tically to propose values corresponding to different exceed- aimed at capturing the nonlinear response of these jacketed
ance probabilities that can be used in practice. components were not developed accordingly.
The data from experiments of jacketed columns presented
NONLINEAR MODELING OF JACKETED in the previous section of this paper were used to compare
COLUMNS the calculated values with measured response to provide
Nonlinear modeling of reinforced concrete frames has confidence in the proposed recommendations developed
become widespread in structural engineering practice with to construct the nonlinear backbone response of jacketed
the availability of software capable to perform these types of columns. Constructing the nonlinear backbone response
analyses efficiently. Although significant efforts have been of jacketed columns requires knowledge of the force
made to develop modeling procedures for existing reinforced developed and its corresponding deformation at various
concrete components, recommendations on how to approach lateral load demands. The procedures followed to determine
modeling of retrofitted frame components are lacking. One these two groups of parameters (force and deformation) at
retrofitting approach is to jacket frame components to various demands were different. Internal forces at different
improve performance, particularly in regions anticipated to demand levels (yield strength, peak strength, and residual
undergo inelastic deformations. Typical reinforcing details strength) were determined using well-established sectional
found in columns in older reinforced concrete frames have models calibrated to closely capture the measured response
been identified as the cause for nonductile column perfor- of jacketed columns in the experiments. The deformations
mance in past earthquakes and related experimental testing. (lateral drift) corresponding to each force level (yield, peak,
The performance of these columns can be improved by jack- drift at residual strength) were determined through statis-

58 ACI Structural Journal/January 2018


tical analysis of available data given the large variation of (maximum of 8300 psi [56 MPa]). Once the confined
drifts observed during the tests and the lack of a robust drift concrete strength was determined, the model developed by
model that captures drift values reliably. The procedures to Mander et al. (1988) was used to construct the complete
determine these force and deformation parameters required uniaxial stress-strain relation for confined concrete, therefore
to construct nonlinear backbone curves of jacketed columns neglecting the further increase in confined concrete strength
are presented in more detail in the following sections. These that results from FRP confinement, particularly at large axial
methods can be easily applied in practice, as their use only strains. It was considered that, for the load levels applied
requires knowledge of the geometric and material properties during testing of the columns included in the database, the
of a given jacketed column. axial strains generated in the compression zone from flexure
would not result in a significant further increase in confined
Lateral load strength of jacketed column sections concrete strength. The effects of confinement from steel or
The sectional model used to determine yield and peak FRP jackets was represented as confinement by an equivalent
force of jacketed columns assumed that behavior is governed amount of transverse reinforcement by equating the lateral
by flexure (no shear failure) after retrofitting and that confining stress from each jacket type to that developed by
shear-moment interaction can be neglected. These assump- transverse reinforcement at a given spacing (Alvarez and
tions were verified by calculating the shear strength of retro- Breña 2014). Confinement efficiency is reduced by arching
fitted columns using the models proposed by Priestley et al. that develops between layers of transverse steel and between
(1994a,b) and Seible et al. (1997) for steel jacketed and FRP longitudinal bars laterally restrained by the corners of hoops
jacketed columns, respectively. At the maximum imposed or ties. These two effects were considered in the model by
displacement in the tests, the ratio of applied load to calcu- decreasing the confining stress efficiency of jackets as appli-
lated shear strength did not exceed 0.86 for FRP-jacketed and cable (refer to Alvarez and Breña [2014] for further details).
0.89 for steel-jacketed columns, respectively. These results Longitudinal reinforcing steel was modeled using an elas-
imply that the jacketed columns included in this study would toplastic material model with strain hardening. The contri-
be expected to develop flexural hinges prior to shear failure, bution of the steel or FRP jackets to flexural strength of the
at the largest displacement demands experienced during cross section was only considered for their effect on concrete
the tests. Flexural yield and flexural strength were there- confinement. For steel jackets, their contribution to increase
fore calculated using a moment-curvature analysis of the the total tension force in the cross section was neglected
retrofitted cross sections. Moment-curvature analysis was because of the potential for slip between the jacket and grout
chosen because of its computational efficiency and accuracy used to fill the concrete-steel gap. FRP jackets are typically
for sections controlled by flexure. After the moment associ- applied with fibers oriented perpendicular to the column axis
ated with each of these two conditions was computed, the so their contribution to increased flexural strength was also
shear force corresponding to yield (Vy) and peak strength neglected. The confinement provided by jacketing is passive,
(VPeak) was determined based on the relationship between activated when the concrete expands due to microcracking.
moment and shear in each test column. The yield and peak Similar to confinement provided by closely spaced hoops,
shear values were compared with the database of jacketed jacket confinement efficiency depends on the properties of
column test results and found to perform well for undam- the jacket (strength and stiffness) and cross-sectional geom-
aged columns prior to retrofitting and for columns with an etry (circular or rectangular column).
axial force ratio (P/(Agfc′)) equal or less than 0.10. The sectional model provided reasonably accurate results
To determine flexural yield and peak moment with the when the original column had not been damaged prior to
sectional model, the column section was discretized using jacketing, and the existing longitudinal reinforcement was
fibers to represent concrete, reinforcing steel, and jacket continuous. The yield and peak strength of columns that
materials. The behavior of each fiber was assumed to be were damaged before jacketing could not be accurately
governed by the uniaxial stress-strain response of each calculated using a moment-curvature approach as described
material, as is commonly done when using fiber section in this paper. In columns containing lapped longitudinal
models. It is worth noting that the behavior of concrete was reinforcement within the plastic hinge region, the maximum
represented by a confined concrete model for concrete in stress in the longitudinal reinforcement was calculated using
compression, assuming that the externally applied jacket a modified version of the equation proposed by Cho and
was effective in confining the concrete throughout the Pincheira (2006), which estimates bar stress as a nonlinear
cross section. The decrease in confinement efficiency of the function of splice length lb in accordance with
jacket, particularly in the case of rectangular cross sections,
was considered as will be described as follows. fs = 1.25(lb/ld)2/3fy ≤ fy (1)
The uniaxial stress-strain behavior of concrete was
modeled including the effects of confinement for concrete where ld is the calculated development length of the bars
in compression and a linear elastic model for concrete in accordance with ACI 318-14, and fy is the nominal
in tension. The strength of concrete confined by steel or yield stress of the bar. Figure 3 compares yield strength
FRP jackets was calculated using the models proposed by (Vy) and peak strength (Vpeak) values determined using the
Priestley et al. (1994a,b) or Lam and Teng (2003a,b), respec- moment-curvature analysis of the column cross section and
tively. The use of these models is suitable for the range of the experimentally determined values of jacketed columns in
concrete strengths found in the jacketed column database the database. For most columns in the database, the discrep-

ACI Structural Journal/January 2018 59


Fig. 3—Comparison between calculated and laboratory jacketed column strength: (a) and (b) shear at yield; and (c) and (d)
shear at peak strength. (Note: 1 kip = 4.45 kN.)
ancy between calculated and measured values at yield and columns are available. The statistical analysis procedure
peak do not exceed 10%, as shown in the figure. Given employed in this process is described in this section.
the simple sectional model used, a difference of 10% in Statistical analysis of drift data for jacketed columns in
calculated and measured shear strength at yield and peak is database—Histograms that show plastic drift data (param-
considered adequate, given the large variations in jacketing eter a) of columns in the database are presented in Fig. 4,
configurations between testing programs and the diverse distinguished only by jacket type (FRP or steel) without
deficiencies that the original columns contained. A sectional separation by cross-sectional shape. Instead of illustrating
model that incorporates the assumptions described in this the distribution of measured deformation data directly,
paper is therefore recommended in practice to estimate yield parameter a (plastic drift capacity) was chosen as a directly
and peak force of jacketed columns. meaningful parameter to represent the nonlinear drift data
in accordance with ASCE/SEI 41-13. Parameter a was
Lateral deformation (drift) of jacketed columns determined for each jacketed column in the database as the
Several deformation components contributed to the total difference between the drift corresponding to a 20% reduc-
drift in jacketed columns in the database. These components, tion in lateral load and the drift at yield defined as described
also common in other reinforced concrete elements, include previously. The data show differences in parameter a
flexural deformation, shear deformation, and rigid body between FRP-jacketed and steel-jacketed columns. FRP-jack-
rotation induced by bar slippage. Any mechanistic model eted columns have a range of parameter a from 0.0073 to
that would capture these effects would be too complex for 0.0882—a mean value of 0.0380 and a ±1 standard devia-
use in design (in contrast to the relatively straightforward tion range of 0.0213 to 0.0547. In contrast, steel-jacketed
sectional model for force levels presented in the previous columns have parameter a-values that range from 0.0149 to
section). Therefore, the database of experimental results was 0.0812, a mean value of 0.0413, and a ±1 standard deviation
used to directly calibrate an empirical model for drift at 80% range of 0.0246 to 0.0580. The drift data and the general
of peak force for columns with low axial force ratios. These behavior observed from the reported hysteresis curves
values could then be used to find nonlinear drift param- also indicated that cross-sectional shape was an important
eter a to construct the backbone nonlinear response curves factor that influenced the deformation behavior of jack-
of jacketed columns. When the axial force ratio was high eted columns. The average plastic deformation capacity of
(>0.60), insufficient data were available and, therefore, the circular sections for both FRP- and steel-jacketed columns
value of parameter a for well-designed unjacketed columns was slightly higher than for rectangular columns, primarily
was adopted. For drift parameter b, the drift at which only due to the higher confining and lap-splice clamping effi-
residual shear capacity remains, values from well-designed ciency of circular jackets over rectangular ones.
unjacketed columns were adopted for both low and high Cross section shape appears to have a meaningful effect
axial force ratios because insufficient test data on jacketed on structural response; therefore, Fig. 4 shows the data histo-

60 ACI Structural Journal/January 2018


Fig. 4—Histograms of parameter a for FRP- and steel-jacketed columns.
grams separated by cross-sectional shape. Although sepa- Table 3—Fitting parameters for Weibull distribution
rating data into groups in this way results in modest sample of parameter a
sizes for each group, it is needed because the mechanics of
Weibull fitting parameters
confinement differ for circular and rectangular columns and,
furthermore, the number of samples is not inconsistent with Jacketing material Cross section shape μ k
the amount of data available for other reinforced concrete Circular 0.0553 2.91
components for which limited laboratory tests have been FRP
Rectangular 0.0380 2.64
conducted. A comparison of parameter a data with three
Circular 0.0471 4.74
typical statistical distributions (lognormal, Weibull, and Steel
Rayleigh) is presented in Fig. 5. These three distributions Rectangular 0.0458 2.36
were chosen because they all satisfy the physical constraint
that parameter a must always have values exceeding zero. reaching residual capacity. This points to a need for further
Although both the Weibull and lognormal distributions testing that extends the testing protocol to the point where
provide good fits to central values of the data, the lognormal the force decays to residual capacity.
distribution has a much heavier upper tail than is indicated
by the data, which would lead to significant probability of RECOMMENDED BACKBONE PARAMETERS FOR
highly unconservative values of drift. Therefore, the Weibull JACKETED COLUMNS
distribution was selected as the appropriate statistical model The ACI 369R-11 contains modeling parameters for
for drift, and fitting parameters for the data shown in Fig. 5 are components of existing frame buildings and is in the process
listed in Table 3. Using this distribution, parameter a values of being updated by using recent test data and statistical
were calculated for a range of probabilities of exceedance studies of available component tests. As noted previously,
between 50 and 95%, as given in Table 4. These data are modeling recommendations for retrofitted components are
provided to allow for selection of design values for param- lacking, so there is a need to provide this information. The
eter a based on exceedance probability in place of the mean approach being followed by ACI 369R is to use mean defor-
value. For both jacketing materials, there is a clear differ- mation values from tests to define modeling parameters. The
ence between parameter a depending on section shape. It is suggested values in this section were developed with this
clear that plastic drift capacity (parameter a) is greater in approach in mind to facilitate future adoption of the tabulated
circular columns than in rectangular columns, as would be values in updates of the ACI 369R and ASCE/SEI 41 docu-
expected given the higher confining and clamping efficiency ments. It is important to note that, due to differences in the
of jackets on circular cross sections. The estimated plastic ability to model force and drift levels for retrofitted columns
drifts that columns can sustain are comparable between jack- and differences in the availability of data for different param-
eted circular columns independent of jacket material. For eters in the backbone curve, a hybrid approach to specifying
rectangular columns, steel-jacketed columns are estimated backbone curve parameters is used herein. For the force
to reach higher plastic drifts than FRP-jacketed columns, levels at yield and peak, a mechanistic sectional model was
as would be expected because of the higher bending stiff- proposed and validated. For drift parameter a, experimental
ness of steel jackets. The comparisons are in line with the data were analyzed and a probabilistic model was fit to the
anticipated differences in behavior of circular and rectan- data. Although herein the mean value is recommended to be
gular columns jacketed with either steel or FRP materials. consistent with current recommended practice, the proba-
It should be noted that although a similar statistical analysis bilistic model is described so that in the future, it will be
of parameter b would be desirable, insufficient data exist possible to specify design values based on desired exceed-
in the open literature to perform such an analysis because ance probability rather than simply using the mean value.
many of the published tests were stopped prior to the column For drift parameter b, recommendations for unretrofitted
columns were adopted in the absence of sufficient data for

ACI Structural Journal/January 2018 61


Fig. 5—Distribution of parameter a and fitted lognormal, Weibull, and Rayleigh statistical distributions.
Table 4—Values of parameter a for selected base, mean data were only computed for parameter a and
probabilities of exceedance for columns tested under low values of axial force (P/(Agfc′)
Parameter a ≤ 0.10). Recommended values of other parameters (b and
c, the residual capacity ratio) are based on the similarities
FRP-jacketed Steel-jacketed
Probability of observed in the behavior of jacketed columns in comparison
exceedance, % Circular Rectangular Circular Rectangular with columns containing reinforcing details representative
95 0.0199 0.0123 0.0251 0.0130 of new design (code conforming columns). The similarities
90 0.0255 0.0161 0.0293 0.0176 in backbone behavior of an originally deficient column that
has been retrofitted using two different jacketing configura-
85 0.0296 0.0190 0.0321 0.0212
tions with the behavior of a similar code-conforming column
80 0.0330 0.0214 0.0343 0.0242 is shown in Fig. 6. The behavior of the jacketed columns
75 0.0360 0.0236 0.0362 0.0270 is remarkably similar to that of columns with well-detailed
70 0.0388 0.0256 0.0379 0.0295 reinforcement, providing support for using modeling param-
eters of these columns for jacketed columns where there are
65 0.0414 0.0275 0.0394 0.0320
no available laboratory data.
60 0.0439 0.0294 0.0408 0.0344 Very few jacketed column tests have been conducted to
55 0.0463 0.0312 0.0422 0.0368 the large displacement demands needed to generate axial
50 0.0487 0.0330 0.0436 0.0392
load failures. One of the few test series of jacketed column
tests conducted to the point of residual lateral load levels is
jacketed columns. This assumption is conservative because the work reported by Ghosh and Sheikh (2007). The tests
jacketing would be expected to increase drift relative to an included reinforced concrete columns containing short
unjacketed column. lap splices retrofitted with FRP jackets. Some of the tests
Based on the statistical analysis of jacketed columns, reported in this research reached rupture of the jacket and
mean values of parameter a for the two cross sections studied lateral load degradation of 90% from the peak strength at
(circular and rectangular) and the two jacket materials included a plastic drift of 11%. These tests demonstrate that plastic
in the database (FRP and steel) were computed (Table 5). drift at strength degradation reaches values that exceed those
These four categories are consistent with the way the data being proposed based on results of well-detailed columns
were organized and statistically studied. Because of the and provides support for the conservative adoption of param-
limitations found in the jacketed column experimental data- eter b values based on well-detailed but unjacketed columns.

62 ACI Structural Journal/January 2018


Table 5—Proposed modeling parameters for FRP- and steel-jacketed columns
Modeling parameters
Section parameters Plastic rotations angle, rad Residual strength ratio
Jacketing material Section shape P/(Agfc′) a b c
≤0.1 0.049 0.060 0.2
Circular
≥0.6 0.010 0.010 0.0
FRP
≤0.1 0.034 0.060 0.2
Rectangular
≥0.6 0.010 0.010 0.0
≤0.1 0.043 0.060 0.2
Circular
≥0.6 0.010 0.010 0.0
Steel
≤0.1 0.040 0.060 0.2
Rectangular
≥0.6 0.010 0.010 0.0

strength of confined concrete can be calculated using models


for confined concrete found in the literature for both FRP-
and steel-confined concrete.
4. In this study, the lack of deformation data for jacketed
columns subjected to large displacements was overcome by
conducting a statistical study of available data to estimate
plastic drift capacity.
5. The mean plastic displacement corresponding to a
lateral load degradation of 20% from peak can be used to
define parameter a to allow construction of backbone curves
that are consistent with ASCE/SEI 41-13.
6. There is a sparsity of test data for jacketed columns
subjected to large displacements that generate a significant
shear force drop from peak (more than 20%). Therefore, in
Fig. 6—Comparison of backbone force-deformation behavior this research, parameter b was based on values that corre-
of code-conforming and two different jacketed columns. spond to the response of columns with code-conforming rein-
forcement details based on the similar hysteretic response
Should further test data become available, it may be possible observed between well-detailed and jacketed columns.
to increase the recommended value of drift parameter b for
jacketed columns. AUTHOR BIOS
ACI member José C. Alvarez is a Visiting Assistant Professor at Quin-
CONCLUSIONS nipiac University, Hamden, CT. He received his PhD from the University
of Massachusetts Amherst, Amherst, MA, in 2017, and his BS in civil and
This paper presents a procedure that can be used in prac- environmental engineering from the University of Puerto Rico, Mayagüez
tice to construct the nonlinear backbone force-deformation Campus, PR. His research interests include seismic design and rehabilita-
response of FRP- or steel-jacketed columns. Backbone tion of new and existing structures.
curves of these retrofitted components can be used to eval- Sergio F. Breña, FACI, is a Professor at the University of Massa-
uate the response of frame structures with selective jack- chusetts Amherst. He received his MS and PhD from the Univer-
eting of columns. Based on the results of this research, the sity of Texas at Austin, Austin, TX. He is a member of ACI Commit-
tees 318-C, Safety, Serviceability and Analysis Structural Concrete
following conclusions and modeling recommendations can Building Code; 369, Seismic Repair and Rehabilitation; 374, Perfor-
be made: mance-Based Seismic Design of Concrete Buildings; 440, Fiber-
1. Section (fiber) models can be used to determine the Reinforced Polymer Reinforcement; and Joint ACI-ASCE Committee 445,
Shear and Torsion. His research interests include the behavior of strength-
lateral load at yield (Vy) and peak (Vpeak) of jacketed columns, ened concrete elements using fiber composites and performance evaluation
as long as the column shear strength exceeds the force of structures through field instrumentation.
required to reach the nominal flexural strength of the section.
Sanjay R. Arwade is a Professor of civil engineering at the University of
2. The contribution of FRP jackets to the flexural strength Massachusetts Amherst. Educated at Princeton and Cornell, his research
of retrofitted sections can be neglected if fibers are oriented interests include probabilistic material and structural mechanics and reli-
in the transverse direction. Steel jacket contribution to flex- ability analysis of structural systems.
ural strength can also be neglected because of potential slip-
ping between jacket and column cross section. ACKNOWLEDGMENTS
The database assembled as part of the research reported in this paper
3. Jackets that extend throughout the plastic hinge region was made possible by funding received from the ACI Concrete Research
of columns can provide concrete confinement and clamping Council. The first author expresses his gratitude to the Northeast Alli-
of short lap splices within this region. The compression ance for Graduate Education and the Professoriate (NEAGEP) at UMass
Amherst for granting him a fellowship to partially support his PhD studies.

ACI Structural Journal/January 2018 63


REFERENCES with Poor Lap-Splice Detailing,” Journal of Bridge Engineering, ASCE,
Aboutaha, R. S.; Engelhardt, M. D.; Jirsa, J. O.; and Kreger, M. E., 1996, V. 10, No. 6, pp. 749-757. doi: 10.1061/(ASCE)1084-0702(2005)10:6(749)
“Retrofit of Concrete Columns with Inadequate Lap Splices by the Use of Haroun, M. A., and Elsanadedy, H. M., 2005b, “Behavior of
Rectangular Steel Jackets,” Earthquake Spectra, V. 12, No. 4, pp. 693-714. Cyclically Loaded Squat Reinforced Concrete Bridge Columns
doi: 10.1193/1.1585906 Upgraded with Advanced Composite-Material Jackets,” Journal of
Aboutaha, R. S.; Engelhardt, M. D.; Jirsa, J. O.; and Kreger, M. E., 1999a, Bridge Engineering, ASCE, V. 10, No. 6, pp. 741-748. doi: 10.1061/
“Rehabilitation of Shear Critical Concrete Columns by Use of Rectangular (ASCE)1084-0702(2005)10:6(741)
Steel Jackets,” ACI Structural Journal, V. 96, No. 1, Jan.-Feb., pp. 68-78. Iacobucci, R. D.; Sheikh, S. A.; and Bayrak, O., 2003, “Retrofit of Square
Aboutaha, R. S.; Engelhardt, M. D.; Jirsa, J. O.; and Kreger, M. E., Concrete Columns with Carbon Fiber-Reinforced Polymer for Seismic
1999b, “Experimental Investigation of Seismic Repair of Lap Splice Fail- Resistance,” ACI Structural Journal, V. 100, No. 6, Nov.-Dec., pp. 785-794.
ures in Damaged Concrete Columns,” ACI Structural Journal, V. 96, No. 2, Lam, L., and Teng, J., 2003a, “Design-Oriented Stress-Strain Model
Mar.-Apr., pp. 297-307. for FRP-Confined Concrete,” Construction and Building Materials, V. 17,
ACI Committee 369, 2011, “Guide for Seismic Rehabilitation of Existing No. 6-7, pp. 471-489. doi: 10.1016/S0950-0618(03)00045-X
Concrete Frame Buildings (ACI 369R-11),” American Concrete Institute, Lam, L., and Teng, J., 2003b, “Design-Oriented Stress-Strain Model
Farmington Hills, MI, 35 pp. for FRP-Confined Concrete in Rectangular Columns,” Journal of Rein-
Alcocer, S. M., and Durán-Hernández, R., 2002, “Seismic Perfor- forced Plastics and Composites, V. 22, No. 13, pp. 1149-1186. doi:
mance of a RC Building with Columns Rehabilitated with Steel Angles 10.1177/0731684403035429
and Straps. Innovations in Design with Emphasis on Seismic, Wind, and Li, Y.-F.; Hwang, J.-S.; Chen, S.-H.; and Hsieh, Y.-M., 2005, “A
Environmental Loading,” Quality Control and Innovations in Materials/ Study of Reinforced Concrete Bridge Columns Retrofitted by Steel
Hot Weather Concreting, pp. 531-552. Jackets,” Zhongguo Gongcheng Xuekan, V. 28, No. 2, pp. 319-328. doi:
Alvarez, J. C., and Breña, S. F., 2014, “Non-Linear Modeling Param- 10.1080/02533839.2005.9670997
eters for Jacketed Columns Used in Seismic Rehabilitation of RC Build- Mander, J. B.; Priestley, M. J. N.; and Park, R., 1988, “Theoret-
ings,” Seismic Assessment of Reinforced Concrete Buildings, SP-297, K. J. ical Stress-Strain Model for Confined Concrete,” Journal of Struc-
Elwood, J. Dragovich, and I. Kim, eds., American Concrete Institute, Farm- tural Engineering, ASCE, V. 114, No. 8, pp. 1804-1826. doi: 10.1061/
ington Hills, MI. (CD-ROM) (ASCE)0733-9445(1988)114:8(1804)
ASCE/SEI 41-13, 2013, “Seismic Rehabilitation of Existing Buildings,” Memon, M. S., and Sheikh, S. A., 2005, “Seismic Resistance of Square
American Society of Civil Engineers, Reston, VA. Concrete Columns Retrofitted with Glass Fiber-Reinforced Polymer,” ACI
Breña, S. F., and Schlick, B. M., 2007, “Hysteretic Behavior of Bridge Structural Journal, V. 102, No. 5, Sept.-Oct., pp. 774-783.
Columns with FRP-Jacketed Lap Splices Designed for Moderate Ductility Ozcan, O.; Binici, B.; and Ozcebe, G., 2010, “Seismic Strengthening of
Enhancement,” Journal of Composites for Construction, ASCE, V. 11, Rectangular Reinforced Concrete Columns Using Fiber Reinforced Poly-
No. 6, pp. 565-574. doi: 10.1061/(ASCE)1090-0268(2007)11:6(565) mers,” Engineering Structures, V. 32, No. 4, pp. 964-973. doi: 10.1016/j.
Chai, Y. H.; Nigel Priestley, M. J. N.; and Seible, F., 1991, “Seismic engstruct.2009.12.021
Retrofit of Circular Bridge Columns for Enhanced Flexural Performance,” Priestley, M. J. N.; Seible, F.; Xiao, Y.; and Verma, R., 1994a, “Steel
ACI Structural Journal, V. 88, No. 5, Sept.-Oct., pp. 572-584. Jacket Retrofitting of Reinforced Concrete Bridge Columns for Enhanced
Cho, J.-Y., and Pincheira, J. A., 2006, “Inelastic Analysis of Reinforced Shear Strength—Part 1: Theoretical Considerations and Test Design,” ACI
Concrete Columns with Short Lap Splices Subjected to Reversed Cyclic Structural Journal, V. 91, No. 4, July-Aug. pp. 394-405.
Loads,” ACI Structural Journal, V. 103, No. 2, Mar.-Apr., pp. 280-290. Priestley, M. J. N.; Seible, F.; Xiao, Y.; and Verma, R., 1994b, “Steel
Galal, K.; Arafa, A.; and Ghobarah, A., 2005, “Retrofit of RC Square Jacket Retrofitting of Reinforced Concrete Bridge Columns for Enhanced
Short Columns,” Engineering Structures, V. 27, No. 5, pp. 801-813. doi: Shear Strength Part 2: Test Results and Comparison with Theory,” ACI
10.1016/j.engstruct.2005.01.003 Structural Journal, V. 91, No. 5, Sept.-Oct., pp. 537-551.
Ghosh, K. K., and Sheikh, S. A., 2007, “Seismic Upgrade With Carbon Seible, F.; Priestley, M. J. N.; Hegemier, G. A.; and Innamorato, D., 1997,
Fiber-Reinforced Polymer of Columns Containing Lap-Spliced Reinforcing “Seismic Retrofit of RC Columns with Continuous Carbon Fiber Jackets,”
Bars,” ACI Structural Journal, V. 104, No. 2, Mar.-Apr., pp. 227-236. Journal of Composites for Construction, ASCE, V. 1, No. 2, pp. 52-62. doi:
Harajli, M. H., and Dagher, F., 2008, “Seismic Strengthening of 10.1061/(ASCE)1090-0268(1997)1:2(52)
Bond-Critical Regions in Rectangular Reinforced Concrete Columns Using Wu, Y.-F.; Wang, L.; and Liu, T., 2008, “Experimental Investigation on
Fiber-Reinforced Polymer Wraps,” ACI Structural Journal, V. 105, No. 1, Seismic Retrofitting of Square RC Columns by Carbon FRP Sheet Confine-
Jan.-Feb., pp. 68-77. ment Combined with Transverse Short Glass FRP Bars in Bored Holes,”
Harajli, M. H., and Rteil, A. A., 2004, “Effect of Confinement Using Journal of Composites for Construction, ASCE, V. 12, No. 1, pp. 53-60.
Fiber-Reinforced Polymer or Fiber-Reinforced Concrete on Seismic doi: 10.1061/(ASCE)1090-0268(2008)12:1(53)
Performance of Gravity Load-Designed Columns,” ACI Structural Journal, Xiao, Y., and Ma, R., 1997, “Seismic Retrofit of RC Circular
V. 101, No. 1, Jan.-Feb., pp. 47-56. Columns Using Prefabricated Composite Jacketing,” Journal of Struc-
Harries, K. A.; Ricles, J. R.; Pessiki, S.; and Sause, R., 2006, “Seismic tural Engineering, ASCE, V. 123, No. 10, pp. 1357-1364. doi: 10.1061/
Retrofit of Lap Splices in Nonductile Square Columns Using Carbon Fiber- (ASCE)0733-9445(1997)123:10(1357)
Reinforced Jackets,” ACI Structural Journal, V. 103, No. 6, Nov.-Dec., Xiao, Y., and Wu, H., 2003, “Retrofit of Reinforced Concrete
pp. 708-716. Columns Using Partially Stiffened Steel Jackets,” Journal of Struc-
Haroun, M. A., and Elsanadedy, H. M., 2005a, “Fiber-Reinforced Plastic tural Engineering, ASCE, V. 129, No. 6, pp. 725-732. doi: 10.1061/
Jackets for Ductility Enhancement of Reinforced Concrete Bridge Columns (ASCE)0733-9445(2003)129:6(725)

64 ACI Structural Journal/January 2018


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 115-S05

Shear-Friction Strength of Low-Rise Walls with 550 MPa


(80 ksi) Reinforcing Bars under Cyclic Loading
by Jang-Woon Baek, Hong-Gun Park, Byung-Soo Lee, and Hyun-Mock Shin

In the thick reinforced concrete walls of nuclear power plants, the The effects of flexural moment and cyclic loading on
use of high-strength reinforcing bars for shear design is necessary shear-friction strength have been investigated in previous
to enhance constructibility and economy. However, low-rise walls studies.13-15 Mattock et al.13 performed a series of push-off
subjected to cyclic lateral loading are susceptible to sliding failure tests for specimens with an interface subjected to flex-
at the construction joints. In the present study, low-rise walls were
ural moment or axial tension. In the Mattock’s tests, the
tested to verify the applicability of 550 MPa (80 ksi) bars to the
yield strength of the reinforcement was 345 to 365 MPa
shear-friction design. The major test parameters were the grade
of reinforcing bars, wall aspect ratio, reinforcement ratio, and (50.0 to 52.9 ksi). The test results showed that flexural
surface condition of the construction joint. The test results showed moment did not decrease the shear-friction strength. The
that the specimens were susceptible to sliding failure and the stress reason was reported in the commentaries of ACI 318-14
of 550 MPa (80 ksi) shear-friction bars was not reached to the (R22.9.4.6)9: flexural compression stress increased the shear-
yield strength. Particularly, the shear-friction strengths under friction strength while flexural tension stress reduced the
cyclic loading were smaller than those under monotonic loading shear-friction strength.
reported in previous studies. The applicability of current design Wood14 proposed a lower-bound shear strength using
methods was evaluated for the shear-friction design of walls with a shear-friction analogy based on the existing test results
550 MPa (80 ksi) bars. of 143 low-rise wall specimens (yield strength of shear-
Keywords: cyclic loading; high-strength reinforcing bars; low-rise shear
friction reinforcement across a shear plane fyv = 280 to
walls; nuclear power plant walls; shear-friction strength. 570 MPa [40.6 to 82.7 ksi]) subjected to monotonic, repeated,
and reversed cyclic loads. The proposed lower-bound shear
INTRODUCTION strength (0.25Av fyv, where Av is area of shear-friction rein-
In the construction of nuclear power plants forcement) is significantly smaller than the shear-friction
(NPPs), many large-diameter reinforcing bars are used for strength of ACI 318 and ACI 349 (μAv fyv; the coefficient of
thick reinforced concrete (RC) walls, which can cause bar shear-friction μ = 0.6 to 1.4).
congestion. Such bar congestion significantly degrades the Orakcal et al.15 tested wall specimens with an intention-
constructability and economy of the structures. Recently, ally weakened joint at the wall-foundation interface, where
to reduce the number of reinforcing bars, the use of high- the concrete cross section was reduced and a part of the
strength 550 MPa (80 ksi) bars for flexural reinforcement vertical reinforcement was discontinued to cause shear-fric-
and shear reinforcement has been studied.1-5 However, tion failure (fyv = 352 MPa [51.0 ksi]). They reported that the
under cyclic loading, squat walls with aspect ratio hw/lw ≤ nominal shear-friction strength of ACI 318 overestimated the
0.5, which are commonly used for nuclear power plants, test strengths (the ratio of the maximum tested strength to
are vulnerable to shear sliding at the construction joint.6,7 the shear-friction strength predicted by ACI 349 [Eq. (A1)]
Thus, in the design of squat walls, frequently, the number of Vtest/Vsf = 0.87 to 0.89), and the strength overestimation was
vertical reinforcing bars is increased to prevent shear sliding. attributed to the effect of cyclic loading. On the other hand,
Currently, the construction joint interfaces are designed Bass et al.16 reported that the effect of cyclic loading on
based on the shear-friction strength specified in Eq. (11-25) of shear-friction strength was negligible (fyv = 420 MPa [60 ksi],
ACI 3498 (or Eq. (22.9.4.2) of ACI 3189). The shear-friction concrete compressive strength fc′ = 18.6 to 34.7 MPa [2.69 to
strength was developed based on experimental and theo- 5.03 ksi]). However, in the test specimens, the shear-friction
retical studies by Birkeland and Birkeland.10 However, the bar ratio ρv was relatively small (= 0.21 to 0.31%). Further, the
majority of the specimens were tested using the push-off number of load cycles (only 10 load cycles) was smaller than
test setup, which differs from the actual loading condition of that of conventional loading protocol.
low-rise walls under earthquake loading in that: 1) flexural Harries et al.17 studied the effect of high-strength rein-
moment as well as shear force is applied to the walls; and forcement. Push-off tests under monotonic loading
2) repeated cyclic loading degrades the shear-friction resis- were performed for shear-friction specimens with normal
tance of the walls. To address such effects, seismic design
code, Eurocode 8,11 specifies stricter requirements than those ACI Structural Journal, V. 115, No. 1, January 2018.
MS No. S-2016-232.R1, doi: 10.14359/51700915, received February 20, 2017, and
of general design code, Eurocode 2.12 On the other hand, in reviewed under Institute publication policies. Copyright © 2018, American Concrete
ACI 3498 (or ACI 3189), no special design requirements are Institute. All rights reserved, including the making of copies unless permission is
obtained from the copyright proprietors. Pertinent discussion including author’s
specified for cyclic loading. closure, if any, will be published ten months from this journal’s date if the discussion
is received within four months of the paper’s print publication.

ACI Structural Journal/January 2018 65


(fyv = 424 to 464 MPa [61.5 to 67.3 ksi]) and high (fyv = 868 friction design. Based on the test results, the relevant design
to 965 MPa [125.9 to 139.9 ksi]) strength bars. The inter- recommendations were provided.
face between old and new concrete was roughened with an
amplitude of at least 6.4 mm (0.25 in.) according to ACI TEST PLAN
3498 or Eurocode 2.12 The test result showed that the high- Major test parameters and specimen details
strength bars did not increase the shear-friction strength. The In NPP walls, because of the high safety requirements,
ratio of the maximum tested strength to the shear-friction the shear reinforcement ratio for the design is close to the
strength predicted by the ACI 349 equations (Vtest/Vsf [Eq. maximum shear reinforcement ratio specified by current
(A1)]) were 1.55 to 2.53 for the specimens with normal- design codes. Thus, most of the specimens tested in the
strength reinforcement, and 0.72 to 1.26 for the specimens present study were designed with the permissible maximum
with high-strength reinforcement. However, in all speci- shear reinforcement ratio (refer to Appendix B for the
mens, the stresses in the both normal-strength and high- maximum permissible ratio). Further, in the case of NPP
strength reinforcing bars did not reach to the yield strength. walls with the permissible maximum shear reinforcement
The test results indicate that under monotonic loading, ratio, uniformly distributed vertical reinforcement, 42 MPa
concrete cohesion significantly contributes to the shear-fric- (6 ksi) compressive strength concrete, a construction joint,
tion strength of walls. and without axial compression, the nominal shear-friction
Kono et al.18 performed push-off tests under reversed strength predicted by the ACI equations becomes smaller
cyclic loading (overall six to eight cycles) to study the effect than the nominal flexural strength, approximately when
of high-strength concrete (44.1 to 112 MPa [6.4 to 16.2 ksi]), M/Vlw ≤ 0.75. Thus, to investigate the shear-sliding failure
high-strength bars (386 to 999 MPa [56.0 to 144.9 ksi]), and strength before flexural yielding, aspect ratios of hw/lw ≤ 0.5
interface roughness (smooth, roughened, and monolithic (M/Vlw ≤ 0.66) were used for the wall specimens.
surface). Similar to the study by Harries et al.,17 even in Thirteen specimens were prepared for cyclic loading tests.
specimens with normal-strength shear-friction bars (386 to Figure 1 shows the dimensions and details of the specimens,
432 MPa [56.0 to 62.6 ksi]), the stresses did not reach and Table 1 presents the design parameters. The dimensions
the yield strength at the measured maximum strength of of walls with hw/lw = 0.5 were 1500 mm (length) x 750 mm
the specimens. Vintzeleou and Tassios19 also reported the (height) x 200 mm (thickness) (59.10 x 29.55 x 7.88 in.). The
strength and stiffness degradation of dowels under cyclic dimensions of walls with hw/lw = 0.33 were 1500 mm (length)
deformation. Such degradation of maximum strength was x 500 mm (height) x 200 mm (thickness) (59.10 x 19.70 x
also found under repeated loading (not reversed loading): 7.88 in.) (Fig. 1 and Table 1). The test parameters were the
Figueira et al.20 reported that the shear-friction strength grade of reinforcing bars, wall aspect ratio, reinforcement
decreased under repeated loading. ratio, surface condition of the construction joint, and addi-
As such, the existing studies reported that the use of tional details (axial load or additional shear-friction bars).
high-strength reinforcement (fyv = 868 to 999 MPa [125.9 to The names of the specimens indicate the test parameters. The
144.9 ksi]) did not increase the shear-friction strength first letters, H and N, refer to higher-strength deformed bars
of walls. However, the range of the yield strength is far (Grade 550 MPa [80 ksi]) and normal-strength deformed bars
beyond the yield strength targeted in the present study, (Grade 420 MPa [60 ksi]), respectively (refer to Table 2). The
and the number of shear-friction tests for walls with Grade numbers 0.5 and 0.33 refer to the aspect ratio of the spec-
550 MPa (80 ksi) reinforcement are limited, particularly for imens. The second letter indicates the horizontal bar ratio:
squat walls (hw/lw ≤ 0.5) with construction joints. Further- M, H, and Q refer to the maximum, half, and quarter of the
more, the effects of cyclic loading and flexural moment on permissible reinforcement ratio of horizontal bars, respec-
shear-friction strength should be meticulously investigated tively. (Refer to Appendix B for the maximum permissible
for the seismic design of squat walls. reinforcement ratio for horizontal bar). The third letter indi-
In the present study, the effect of high-strength bars on the cates the surface condition of the construction joint: U, R, and
shear-friction strength of low-rise walls subjected to cyclic G refers to untreated surface, roughened surface, and grooved
loading was studied. The test parameters were the bar grades surface, respectively. The last letter(s) following the dash (-)
(Grade 420 and 550 MPa [60 and 80 ksi] bars), aspect ratio indicate the additional details: C for axial compression load
hw/lw (0.5 and 0.33), reinforcement ratio, surface condition (0.06 to 0.07Acfc′, where Ac is area of wall cross section),
of construction joint, presence of axial compression, and use AS for additional short shear-friction bars (bar length =
of additional shear-friction bars. 160 mm [6.30 in.]), and AL for additional long shear-friction
bars (300 mm [11.82 in.]). For example, H0.5MU-C indicates
RESEARCH SIGNIFICANCE a specimen with Grade 550 MPa (80 ksi) shear bars, an aspect
In the thick RC walls with small aspect ratios of NPPs, ratio of 0.5, maximum shear reinforcement ratio, untreated
shear sliding failure along the construction joints is the surface in the construction joint, and axial compression
major failure mode under earthquake loading. Thus, for the loading (0.06 to 0.07Ac fc′).
use of high-strength bars for NPP walls, it is crucial to verify For Specimen H0.5MU, Grade 550 MPa (80 ksi) rein-
the effect of high-strength bars on the shear-friction strength. forcing bars and aspect ratio hw/lw = 0.5 were used. For
The present study experimentally evaluated the validity horizontal and vertical reinforcement, two layers of HD13
of high-strength bars (Grade 550 MPa [80 ksi]) for shear bars (actual yield strength fyh and fyv = 667 MPa [96.7 ksi])
were used (Fig. 1(b), Table 2). The horizontal bar ratio ρh

66 ACI Structural Journal/January 2018


Fig. 1—Dimensions and reinforcement details of test specimens. (Note: 1 MPa = 0.145 ksi; 1 mm = 0.0394 in.)
was 0.68% (Table 1), which corresponds to the maximum flexural strength was increased by the axial force, the
shear bar ratio specified in ACI 349 (using Vs [Eq. (B3)] = number of flexural bars was decreased when compared to
Vsmax [Eq. (B4)]). The vertical bar ratio ρv was 1.01%. The that of H0.5MU (ρv = 0.68%, ρh = 0.68%). However, the
nominal flexural strength Vf (that is, horizontal force corre- nominal shear-friction strength Vsf was also increased by the
sponding to the flexural moment capacity) was designed to compression and was close to the nominal flexural strength
be equal to the nominal shear strength Vn (Eq. (B1) to (B3)). Vf (Vsf/Vf = 1.02).
A construction joint without surface treatment was located In Specimen H0.5MG-C, to prevent sliding failure, the
at the interface between the wall and foundation slab. Thus, wall-foundation interface was intentionally moved approx-
the friction coefficient of the nominal shear-friction strength imately 50 mm (1.97 in.) to the inside of the foundation
μ was 0.6 (Eq. (A1)). All the vertical bars were assumed to (Fig. 1(a)). To achieve this purpose, pocket metals were
contribute to the nominal shear-friction strength Vsf, which installed at the wall-foundation interface. After pouring
was close to the nominal flexural strength Vf (or the nominal concrete for the foundation, a groove was formed by the
shear strength Vn) (Vsf/Vf = 0.97). pocket metals. Thus, the wall-foundation interface in
In Specimens H0.5HU and H0.5QU, to investigate the effect H0.5MG-C was regarded as a monolithic joint: the friction
of the bar ratio, smaller bar ratios were used (Fig. 1(b)): ρh = coefficient μ = 1.4. Thus, the nominal shear-friction strength
0.38%, ρv = 0.60% for H0.5HU and ρh = 0.19%, ρv = 0.43% was significantly increased to 2921 kN (657.2 kip), but was
for H0.5QU (Table 1). The nominal shear-friction strength limited to 1919 kN (431.8 kip) (Eq. (A2a), ACI 318) or
Vsf was smaller than the nominal flexural strength Vf (or the 1650 kN (371.3 kip) (Eq. (A2b), ACI 349) by the permissible
nominal shear strength Vn) (H0.5HU: Vsf/Vf = 0.82, H0.5QU: maximum value specified by ACI 318 or ACI 349 (Vsf/Vf =
Vsf/Vf = 0.76). 1.49 or 1.28), respectively. The other details of H0.5MG-C
In Specimen H0.5MU-C, the effect of compressive axial were the same as those of H0.5MU-C (Fig. 1(b)).
force (0.07Ac fc′) was studied (Fig. 1(b)). As the nominal

ACI Structural Journal/January 2018 67


Table 1—Design parameters of test specimens
Additional shear-
Wall
friction reinforcing
aspect
Wall concrete Horizontal reinforcing bar Vertical reinforcing bar bar
ratio
hw/lw fc′, Number and ρhfyh, Number
Specimens (M/Vlw) MPa P, kN Age Vs/Vsmax type ρh, % MPa Number and type ρv, % and type ρsf, % Vsf /Vf
H0.5MU 42.1 0 21 1.04 8-HD13 0.68 4.51 24-HD13 1.01 — — 0.97
H0.5HU 42.1 0 20 0.55 8-HD10 0.38 2.38 12-HD13/4-HD10 0.60 — — 0.82
0.5
H0.5QU 42.1 0 20 0.27 4-HD13 0.19 1.19 8-HD13/4-HD10 0.43 — — 0.76
(0.66)
H0.5MU-C 42.1 882 17 1.04 8-HD13 0.68 4.51 16-HD13 0.68 — — 1.02
H0.5MG-C 38.7 735 51 1.09 8-HD13 0.68 4.51 16-HD13 0.68 — — 1.28
H0.33MU 42.1 0 24 1.03 10-HD10 0.71 4.46 18-HD13 0.76 — — 0.68
H0.33HU 42.1 0 22 0.62 6-HD10 0.43 2.67 12-HD13 0.51 — — 0.64
H0.33MU-AS 42.1 0 27 1.03 10-HD10 0.71 4.46 18-HD13 0.76 16-HD10 0.38 1.00
H0.33HU-AS 0.33 42.1 0 23 0.62 6-HD10 0.43 2.67 12-HD13 0.51 10-HD10 0.24 0.93
H0.33MU-AL (0.50) 45.5 0 21 0.99 10-HD10 0.71 4.46 18-HD13 0.76 16-HD10 0.38 1.00
N0.33MU 44.5 0 19 1.16 8-D13 1.01 5.17 16-D16 1.06 — — 0.67
N0.33MU-AL 44.5 0 19 1.16 8-D13 1.01 5.17 16-D16 1.06 14-D13 0.59 1.07
H0.33MR 44.0 0 20 1.01 10-HD10 0.71 4.46 18-HD13 0.76 — — 1.13

Notes: M/Vlw is shear span ratio; fc′ is compressive strength; P is axial compression; Vs is contribution of shear reinforcement (Eq. (B3)); Vsmax is maximum shear strength contrib-
uted by shear reinforcement in ACI 349 (Eq. (B4)); Vsf is shear-friction strength prediction (Eq. (A1) and Eq. (A2b)); Vf is flexural strength prediction; HD is Grade 550 MPa
(80 ksi) deformed bars; D is Grade 420 MPa (60 ksi) deformed bars (refer to Table 2); 1 kN = 0.225 kip; 1 MPa = 0.145 ksi.

Table 2—Properties of reinforcement U-type bars were respectively placed at the wall-foundation
interface (Fig. 1(c) and 1(d)). Consequently, the nominal
Type HD10 HD13 D13 D16
shear-friction strength was close to the nominal flex-
Grade 550 MPa (80 ksi) 420 MPa (60 ksi) ural strength (Vsf/Vf = 1.00 for H0.33MU-AS and 0.93 for
9.5 mm 12.7 mm 12.7 mm 15.9 mm H0.33HU-AS). For the U-type bars, HD10 (yield strength
Bar diameter
(No. 3) (No.4) (No.4) (No.5) of additional shear-friction reinforcement fysf = 625 MPa
Yield strength 625 667 510 470 [90.6 ksi]) was used (Table 2). The U-type shear-friction
fy, MPa (ksi) (90.6) (96.7) (74.0) (68.2) reinforcement was assumed to not contribute to the flexural
Notes: HD is Grade 550 MPa (80 ksi) deformed bars; D is Grade 420 MPa (60 ksi)
strength Vf. The length of the U-type bars from the wall-foun-
deformed bars; 1 mm = 0.0394 in. dation interface to the end of the hook was 160 mm (6.30 in.)
to satisfy the development length of the hook (Fig. 1(c)).
In the case of walls with hw/lw = 0.5, except H0.5HU and In Specimen H0.33MU-AL, the development length of the
H0.5QU, the nominal shear-friction strength Vsf was close U-type bars from the wall-foundation interface to the end
to the nominal flexural strength Vf. On the other hand, in the of the hook was increased to 300 mm (11.82 in.)—approxi-
case of walls with hw/lw = 0.33, the nominal flexural strength mately 30-bar diameter. For the development length of addi-
Vf for a given vertical reinforcing bar area was increased due tional shear-friction reinforcement (straight reinforcing bars
to the smaller wall height. Therefore, the nominal shear-friction without hook), Wasiewicz7 recommended the greater value
strength was smaller than the nominal flexural strength. of 15 bar diameter and wall thickness. The other details of
In Specimen H0.33MU, the ratio of the nominal H0.33MU-AL were the same as those of H0.33MU-AS.
shear-friction strength to the nominal flexural strength was To directly compare the effects of Grade 550 and 420 MPa
significantly small: Vsf/Vf = 0.68. Like H0.5MU, the permis- (80 and 60 ksi) bars, in Specimen N0.33MU, D13 and D16
sible maximum shear bar ratio was used for shear reinforce- bars [Grade 420 MPa (60 ksi)] were used for the horizontal
ment (Table 1, ρh = 0.71%, ρv = 0.76%). For horizontal and bars and vertical bars, respectively (Fig. 1(d), Tables 1 and 2).
vertical reinforcement, HD10 (fyh = 625 MPa [90.6 ksi]) and The horizontal bar ratio was increased to 1.01% as the yield
HD13 (fyv = 667 MPa [96.7 ksi]) were used, respectively. strength of bars was decreased. The effective bar strength
In H0.33HU, to study the effect of bar ratios, the bar ratios ρhfyh = 5.17 MPa (0.75 ksi) was slightly greater than that of
were decreased (ρh = 0.43%, ρv = 0.51%). H0.33MU (ρhfyh = 4.46 MPa [0.65 ksi]). The vertical bar ratio
In Specimens H0.33MU-AS and H0.33HU-AS, addi- was ρv = 1.06%. In N0.33MU-AL, U-type D13 bars (fysf =
tional shear-friction bars were placed at the wall-foundation 510 MPa [74.0 ksi], ρsf = 0.59%) were additionally placed at
interface so that the nominal shear-friction strength Vsf was the wall-foundation interface. Similar to H0.33MU-AL, the
close to the nominal flexural strength Vf. In H0.33MU-AS length of the U-type bars from the wall-foundation interface
and H0.33HU-AS, to increase only the shear-friction to the end of hook was 300 mm (11.82 in.). The other details
strength rather than the flexural strength, eight and five short of N0.33MU-AL were the same as those of N0.33MU.

68 ACI Structural Journal/January 2018


Table 3—Mixture design of concrete
Unit weight, kgf/m3
Maximum
Compressive strength fc′ w/c, % W C FS S FA CA SP Slump, mm aggregate size, mm
42 MPa 33.3 161 340 58 85 749 956 4.1 120 25

Notes: W is water; C is cement; FS is fly ash; S is blast-furnace slag; FA is fine aggregate; CA is coarse aggregate; SP is superplasticizer (high-range water-reducing admixture);
1 mm = 0.0394 in.; 1 m = 3.281 ft; 1 kgf = 2.204 lbf.

Fig. 3—Lateral loading protocol. (Note: 1 mm = 0.0394 in.)


Fig. 2—Test setup. (Note: 1 mm = 0.0394 in.). ment transducers (LVDTs) for the measurement of lateral
displacements, flexural deformations, shear deformations,
In Specimen H0.33MR, to satisfy the “roughened surface”
and sliding at the wall base. Figures 1(a), 1(c), and 1(e) show
specified by ACI 3498 and Eurocode 2,12 the interface
the strain gauges for the reinforcing bars.
was roughened with surface treatment of 6 mm (0.24 in.)
(Fig. 1(d) and (e)). For the roughened surface, the fric-
TEST RESULTS
tion coefficient μ was increased to 1.0. The other details of
Load-displacement relationships
H0.33MR were the same as those of H0.33MU.
Figure 4 shows the lateral load-displacement (lateral drift
In the test specimens, the concrete compressive strength
ratio) relationships of the test specimens. The lateral displace-
was 42.1 to 45.5 MPa (6.10 to 6.60 ksi), except for 38.7 MPa
ment indicates the net displacement (L1-L2 in Fig. 2) excluding
(5.61 ksi) of H0.5MG-C (Table 1). The ages of the concrete
sliding of the wall base (L2 in Fig. 2). Figure 4 also shows
at the time of testing are shown in Table 1. Table 3 presents
the nominal shear strength Vn, nominal flexural strength Vf,
the mixture design of concrete. In the small-scale test spec-
and nominal shear-friction strength Vsf predicted by ACI 349
imens, the wall depth was not sufficiently large to provide
(ACI 318). In the case of H0.33MU-AS, H0.33HU-AS,
good anchorage of the horizontal bars. Thus, 90-degree
H0.33MU-AL, and N0.33MU-AL, the nominal flexural
hooks were used at the ends of the horizontal bars. All the
strength Vf was calculated assuming that the additional
vertical bars and U-type shear-friction bars satisfied the
shear-friction bars do not contribute to the flexural strength.
anchorage length in the foundation slab. The concrete clear
Due to the relatively short length, the flexural contribution of
cover was 20 mm (0.79 in.).
the additional vertical reinforcing bars is limited only to the
wall bottom, and disappears in the area above the wall bottom.
Test procedure and instrumentation
In the calculation of the nominal shear-friction strength,
A cyclic lateral load was applied using the test setup
the limitation of bar yielding strength (420 MPa [60 ksi])
shown in Fig. 2. In the case of H0.5MU-C and H0.5MG-C,
specified in ACI 349 was not used. In all the specimens
an axial compressive load of approximately 0.06 to 0.07Ac fc′
except H0.5MG-C, H0.33MU-AL, and N0.33MU-AL, the
was applied to the top of the wall (882 kN [198.5 kip] in
maximum tested strength Vtest was not reached to the nominal
H0.5MU-C (42.1 MPa [6.10 ksi]), and 735 kN [165.4 kip]
flexural strength Vf. This result indicates that the maximum
in H0.5MG-C (38.1 MPa [5.52 ksi])). The level of the axial
tested strength Vtest was determined by shear sliding failure
compressive force was maintained during cyclic lateral
before flexural yielding (Fig. 4(a) to 4(d), 4(f) to 4(j), and
loading. Generally, the actual level of axial load ratio for
4(m)). On the other hand, in H0.5MG-C showing flexural
NPP walls ranged from 0.05 to 0.20Ac fc′. The axial load
yielding, the test strength was close to the flexural strength
level of 0.06 to 0.07Ac fc′ was determined: 1) to consider
(Fig. 4(e)). In H0.33MU-AL and N0.33MU-AL, with addi-
limitation of the axial load capacity of the testing actuator;
tional shear-friction reinforcing bars of 300 mm (11.82 in.), the
and 2) to compare the test results with the results of previous
test strength was reached to the nominal flexural strength
studies4,21 by maintaining the same axial load ratio.
(Fig. 4(k) and 4(l)). However, it should be noted that the
The lateral loading protocol in Fig. 3 followed the
nominal flexural strength was calculated neglecting the
“Acceptance Criteria for Special Precast Concrete Struc-
contributions of the additional vertical reinforcing bars.
tural Walls.”22 Figure 2 shows the linear variable displace-
Thus, the actual flexural strength could be greater than

ACI Structural Journal/January 2018 69


Fig. 4—Lateral load-displacement relationships of specimens. (Note: 1 mm = 0.0394 in.; 1 kN = 0.225 kip.)
the nominal flexural strength. The flexural strength Vf was strength of shear friction is limited to 420 MPa [60 ksi] (refer
calculated by sectional analysis considering the effect of to the section, “Strains of reinforcing bars”).
axial load. In H0.5HU and H0.5QU (Fig. 4(b) and 4(c)) with smaller
In Specimen H0.5MU (Fig. 4(a)) with HD13 bars (fyh and bar ratios, the maximum tested strength Vtest decreased:
fyv = 667 MPa [96.7 ksi], ρh = 0.68%, ρv = 1.01%), the peak in the case of H0.5HU, +450 and –544 kN (+101.3 and
strength occurred at the drift ratio of +0.45 and –0.60%. The –122.4 kip) at the drift ratios of +0.35 and –0.60%, and in the
maximum tested strength Vtest was +744 and –831 kN (+167.4 case of H0.5QU, +441 and –406 kN (+99.2 and –91.4 kip)
and –187.0 kip) in the positive and negative loading directions, at the drift ratios of +0.28 and –0.60%. The maximum tested
respectively, which was only 65% of the nominal shear-fric- strength was approximately proportional to the vertical rein-
tion strength predicted by the ACI 349 (or ACI 318) equation; forcing bar ratio. However, the ratio of the maximum tested
the nominal shear-friction strength significantly overestimated strength to the nominal shear-friction strength by ACI 349
the tested strength. This is mainly because the yield strength (Vtest/Vsf) was increased to 0.70 and 0.83 in H0.5HU and
of Grade 550 MPa (80 ksi) reinforcement was assumed to be H0.5QU, respectively. This result indicates that the nominal
valid for the nominal shear-friction strength. However, in fact, shear-friction strength predicted by ACI 349 and ACI 318
in the current ACI 349 (or ACI 318) equation, the design yield equations is overestimated in the case of higher bar ratios.

70 ACI Structural Journal/January 2018


Table 4—Summary of test results
Ratio of test strengths
Test results to predictions Estimated bar yield stress
Specimens Vtest,s, kN Vtest,g, kN Vtest, kN Drift ratio at Vtest, % Failure mode Vtest,s/Vsf Vtest,s/Vf fsf [= (Vtest,s – μP)/(μAvall)], MPa
H0.5MU 744 831 787 +0.45/–0.6 0.61 0.59 408
H0.5HU 450 544 497 +0.35/–0.6 0.63 0.52 370
SF
H0.5QU 406 442 423 +0.28/–0.6 0.79 0.60 445
H0.5MU-C 1087 1089 1087 +1.0/–1.0 0.81 0.83 458
H0.5MG-C 1368 1440 1404 +1.0/–1.0 CC+DT 0.83 1.06 —
H0.33MU 575 718 646 +0.35/–0.45 0.63 0.43 420
SF
H0.33HU 309 535 421 +0.45/–0.2 0.51 0.33 339
H0.33MU-AS 1031 1060 1045 +1.0/–0.6 0.77 0.77 502
SW
H0.33HU-AS 709 743 726 +1.0/–0.6 0.81 0.75 497
H0.33MU-AL 1344 1492 1418 +1.25/–1.25 1.00 1.00 655
SF
N0.33MU 969 1040 1004 +0.45/–0.6 1.08 0.72 508
N0.33MU-AL 1321 1531 1426 +1.0/–1.0 SW 0.92 0.99 444
H0.33MR 1132 1351 1242 +1.0/–1.0 SF 0.74 0.84 497

Notes: Vtest,s, Vtest,g, and Vtest are smaller, greater, and average values of measured maximum loads in positive and negative loading directions, respectively; Vsf is shear-friction
strength predictions by ACI 349 (Eq. (A1)); Vf is flexural strength predictions; SF is sliding failure at wall-foundation interface; CC is concrete crushing failure; DT is diagonal
tension failure; SW is sliding failure at midheight of wall; 1 kN = 0.225 kip; 1 MPa = 0.145 ksi.

In H0.5MU-C (Fig. 4(d)), subjected to axial compression, strength may be caused by an unexpected loading condition
the maximum tested strength Vtest and the corresponding such as torsion, which occurs under the pull-push lateral
lateral drift significantly increased: +1088 and –1086 kN loading. Similar to the specimens with hw/lw = 0.5 (H0.5MU,
(+244.8 and –244.4 kip) at the drift ratios of +1.0 and H0.5HU, and H0.5QU), the maximum tested strength was
–1.0%, respectively. Assuming that the vertical reinforcing proportional to the vertical reinforcing bar ratio, which indi-
bar ratio of H0.5MU-C (ρv = 0.68%) was identical to that cates that the shear-friction strengths were not affected by the
of H0.5HU (ρv = 0.60%), the shear-friction coefficient μ wall aspect ratios between hw/lw = 0.5 and 0.33.
for axial compression can be estimated as approximately In N0.33MU (Fig. 4(h)) with D13 and D16 bars (fyh =
0.67 [= (1087 – 497)kN/882 kN)], which coincides with the 510 MPa [74.0 ksi], ρh = 1.01%, fyv = 470 MPa [68.2 ksi],
coefficient for the smooth surface (not intentionally rough- ρv = 1.06%), unlike the walls with HD10 and HD13 bars,
ened) specified in the ACI shear-friction equation (refer to the maximum tested strength was reached to the nominal
Appendix A). Nevertheless, the maximum tested strength shear-friction strength Vsf: +1039 and –968 kN (+233.8 and
was not reached to the shear-friction strength: Vtest/Vsf = 0.81. –217.8 kip) at the drift ratios of +0.45 and –0.60%, respec-
In H0.5MG-C (Fig. 4(e)), with a groove construction joint tively. After the peak load, as the lateral drift increased, the
and axial compression loading, the maximum tested strength load-carrying capacity gradually decreased.
Vtest exceeded the nominal flexural strength Vf: +1368 and In H0.33MU-AS and H0.33HU-AS (Fig. 4(i) and 4(j))
–1440 kN (+307.8 and –324.0 kip) at the drift ratios of +1.0 using HD10 additional shear-friction bars (fysf = 625 MPa
and –1.0%, showing ductile behavior until the drift ratios of [90.6 ksi], ρsf = 0.38% in H0.33MU-AS, and ρsf = 0.24%
+2.0 and –2.0%, respectively. in H0.33HU-AS), the maximum tested strength Vtest and
In H0.33MU (Fig. 4(f)) with aspect ratio hw/lw = 0.33 (fyh = the corresponding lateral drift increased: in H0.33MU-AS
625 MPa [90.6 ksi], ρh = 0.71%, fyv = 667 MPa [96.7 ksi], ρv = (maximum shear bar ratio), +1060 and –1031 kN (+238.5
0.76%), the maximum tested strength was smaller than that of and –232.0 kip) at the drift ratios of +1.0 and –0.60%, and in
H0.5MU: +574 and –718 kN (+129.2 and –161.6 kip) at the H0.33HU-AS (smaller shear bar ratio), +708 and –743 kN
drift ratios of +0.35 and –0.45%, respectively. In H0.33HU (+159.3 and –167.2 kip) at the drift ratios of +1.0 and –0.60%,
(Fig. 4(g)), with half of the permissible maximum bar ratio, respectively. However, the maximum tested strength did not
the maximum tested strength Vtest decreased: +309 and reach the nominal shear-friction strength predicted by the
–534 kN (+69.5 and –120.2 kip) at the drift ratios of +0.45 ACI 349 equation: Vtest/Vsf = 0.78 and 0.83 in H0.33MU-AS
and –0.20%, respectively. The ratios of the maximum tested and H0.33HU-AS, respectively. On the other hand, in
strength to the nominal shear-friction strength predicted by H0.33MU-AL (fysf = 625 MPa [90.6 ksi], ρsf = 0.38%) and
ACI 349 equation were Vtest/Vsf = 0.71 and 0.69 in H0.33MU N0.33MU-AL (fysf = 510 MPa [74.0 ksi], ρsf = 0.59%) with
and H0.33HU, respectively. In H0.33HU, the test strength the longer shear-friction bars (300 mm [11.82 in.]) (Fig. 4(k)
(309 kN [69.4 kip]) in the positive direction was excessively and 4(l)), the maximum tested strength was reached to the
less than the test strength (534 kN [120.1 kip]) in the negative nominal shear-friction strength: for H0.33MU-AL, Vtest/Vsf  =
direction, though the other specimens also showed a similar 1.06 (Vtest = +1491 and –1344 kN [+335.5 and –302.4 kip]
trend (refer to Table 4). The excessive difference in the test at +1.25 and 1.25%, respectively), and for N0.33MU-AL,

ACI Structural Journal/January 2018 71


Fig. 5—Damage modes of specimens at end of test.
Vtest/Vsf = 0.99 (Vtest = +1530 and –1321 kN [+344.3 and of the wall height (167 mm [6.58 in.] from the wall-foundation
–297.2 kip] at +1.0 and –1.0%, respectively). This result interface) (Fig. 5(d), SW mode). Similarly, in N0.33MU-AL,
indicates that the development length of the shear friction with additional D13 shear-friction bars of 300 mm (11.82 in.)
reinforcing bars should be greater than the ordinary devel- anchorage length, the failure surface was formed along three-
opment length specified by current design codes. This is fifths of the wall height (Fig. 5(e)¸ SW mode). On the other
because at the bottom of walls subjected to cyclic loading, hand, in H0.33MU-AL with additional HD10 shear-friction
wide and severe flexure-shear cracks degrade the bond bars of 300 mm (11.82 in.) anchorage length, the failure
strength of the shear friction reinforcing bars. In the present surface occurred at the wall-foundation interface (SF mode).
study, development length of a 30-bar diameter is recom- In H0.5MG-C, with a groove construction joint, sliding
mended for the U-type Grade 550 MPa [80 ksi] shear-fric- failure did not occur, and a mixed failure mode of shear
tion reinforcement. and flexure occurred: crushing in the compression zone and
In H0.33MR (Fig. 4(m)), with a roughened surface, the diagonal tension cracking in the tension zone (Fig. 5(f),
maximum tested strength was +1350 and –1132 kN (+303.8 CC+DT mode).
and –254.7 kip), which was 1.92 times that of H0.33MU Figure 6 shows the crack patterns of H0.33MU,
with the smooth surface. Nevertheless, the maximum tested H0.33MU-AS, and H0.5MG-C, which represent the failure
strength did not reach the nominal shear-friction strength: modes of the wall-foundation interface (SF mode), wall
Vtest/Vsf = 0.82 (m = 1.0). interface (SW mode), and combined flexure and shear
(CC+DT mode), respectively. In the SF mode specimens
Failure modes (Fig. 6(a)), at the drift ratio of 0.2%, vertical splitting tensile
Figure 5 shows the damage modes of the specimens at the cracks occurred at the wall-foundation interface at the ends
end of the tests. Table 4 summarizes the failure modes. In of the wall section. Although diagonal cracking occurred in
H0.5MU and H0.5MU-C, although the design shear-friction the web, the number and width of the diagonal cracks did not
strength was the same as the design flexural strength, sliding increase after 0.6% drift ratio. After the drift ratio of 1.0%,
failure occurred at the wall-foundation interface (Fig. 5(a), SF the extensive vertical splitting tensile cracks led to spalling
mode). The failure mode was the same as that of H0.5HU, of the concrete in the wall ends, and additional vertical split-
H0.5QU, H0.33MU (Fig. 5(b)), H0.33HU, and N0.33MU ting tensile cracks occurred in the web of the wall. Spalling
(Fig. 5(c)), which were intentionally designed for sliding of the concrete at the wall-foundation interface became
failure (Table 1). The test results showed that the failure mode severe, and ultimately, sliding failure occurred along the
of the specimens was determined by shear sliding before flex- wall-foundation interface (Fig. 5(b)). In H0.5MU, with an
ural yielding irrespective of the vertical reinforcing bar area, aspect ratio of 0.5, vertical splitting tensile cracks occurred
because the increase of vertical reinforcing bars increased earlier (0.2% drift ratio) in the web as well as at the ends of
the flexural capacity as well as shear-friction capacity. In the wall. In H0.5HU, H0.5QU, and H0.33HU, with smaller
H0.33MR, with the roughened construction joint (μ = 1.0), bar ratios, sliding failure occurred earlier. The number of
sliding failure occurred at the wall-foundation interface, cracks was fewer than that of H0.5MU, H0.33MU, and
although the design shear-friction strength was greater than N0.33MU, which had greater bar ratios. In H0.33MR, with
the design flexural strength (Vsf /Vf = 1.13, Table 1). the roughened construction joint (μ = 1.0), the failure mode
In Specimens H0.33MU-AS, H0.33HU-AS, H0.33MU-AL, was the same as that of H0.33MU, except that extensive
and N0.33MU-AL, despite the use of additional shear-fric- diagonal cracking also occurred.
tion bars, sliding failure occurred. In H0.33MU-AS and In the SW mode specimens (Fig. 6(b)), diagonal cracking
H0.33HU-AS, the sliding failure surface was formed along occurred at the beginning of the test and spread across
the end of the shear-friction bars, which was located at a third the entire wall panel until the failure of the wall. Flexural

72 ACI Structural Journal/January 2018


Fig. 6—Crack propagations of specimens.
cracking was initiated at the height where the additional shear-friction bars yield as the peak load is attained, similar
shear-friction bars were terminated (160 mm [6.30 in.] from to the current ACI 349 provisions for bars with yield strength
the wall-foundation interface in the case of H0.33MU-AS). up to 420 MPa (60 ksi). In H0.5HU, H0.5QU, H0.5MU-C,
As the load increased, the width of the flexural cracks H0.33MU, and H0.33HU, with HD13 and/or HD10 bars
significantly increased. At the maximum load, a new hori- and without additional shear-friction bars, the strains of the
zontal crack was formed across the midheight of the wall vertical bars were similar to those of H0.5MU.
and sliding failure occurred along the horizontal crack. In On the other hand, in N0.33MU (Fig. 7(c) and 7(d)), with
the case of H0.33MU-AL, like other SW mode specimens, D16 and D13 bars, bar yielding occurred in the web and
flexural cracking occurred at the end of the shear-friction compression zone as well as in the tension zone at 1.0Vtest. In
bars. However, sliding occurred at the wall-foundation inter- N0.33MU-AL and H0.33MU-AL, with the longer (300 mm
face (SF mode). [11.82 in.]) additional shear-friction bars (Fig. 7(e) to 7(h)),
In the CC+DT mode specimen (Fig. 6(c)), similar to the the strains of vertical bars were similar to those of N0.33MU.
SW mode specimens, diagonal cracking gradually propa- Thus, in the cases of N0.33MU-AL and H0.33MU-AL, with
gated across the entire wall panel. As the drift ratio increased, additional shear-friction bars, the assumption of ACI 349
the width of the diagonal cracks increased, and after flexural shear-friction method was valid, and the maximum tested
cracking, spalling of concrete and buckling of the vertical strength was reached to the design shear-friction strength:
bars occurred at the end of the walls. Ultimately, concrete Vtest/Vsf = 0.99 and 1.06 in N0.33MU-AL and H0.33MU-AL,
crushing and diagonal tensile cracking occurred in the respectively (Table 4). However, in the case of H0.33MU-AS,
compression zone and the tension zone (Fig. 5(f)), respec- with the shorter additional shear-friction bars (160 mm
tively. Unlike the SF mode specimens, vertical splitting did [6.30 in.]), the maximum tested strength did not reach the
not occur during the test. design shear-friction strength despite the use of additional
shear-friction bars: Vtest/Vsf = 0.83. Similar to H0.5MU, the
Strains of reinforcing bars strains of the vertical bars were linearly distributed, and bar
Figure 7 shows the strain distributions of the vertical bars yielding did not occur in the vertical bars or the additional
along the wall length at 0.5Vtest, 0.75Vtest, and 1.0Vtest, where bars (Fig. 7(i) and 7(j)).
Vtest indicates the maximum tested strength. The locations of Figure 8 shows the strain distributions of the horizontal
the bars and strain gauges are shown in Fig. 7. The strains bars along the wall height (Fig. 1(a)) at 0.5Vtest, 0.75Vtest, and
that were measured near the failure interface are presented. 1.0Vtest. In H0.5MG-C (Fig. 8(a)), without sliding failure, the
In all specimens, the strains of the vertical bars and additional strains of the horizontal bars were greater than those of the
shear-friction bars were approximately linearly distributed specimens with sliding failure (Fig. 8(b) and 8(c)). However,
at 0.5 to 0.75Vtest because of the flexural action (Fig. 7). yielding of the horizontal bars did not occur because of: 1)
At 1.0Vtest in H0.5MU (Fig. 7(a) and 7(b)), the strains of the smaller contribution of horizontal bars in the low aspect
the vertical bars were still linear, and bar yielding did not ratio (hw/lw = 0.5); and 2) the higher-yield strain (≈0.0033) of
occur. This test result contradicts the assumption that all the high-strength bars.

ACI Structural Journal/January 2018 73


Fig. 7—Measured strains of vertical bars and shear-friction bars in specimens. (Note: 1 mm = 0.0394 in.; 1 kN = 0.225 kip.)
EVALUATION OF DESIGN STRENGTH OF
SHEAR-FRICTION BARS
Figures 9(a) to 9(d) show the ratios of the maximum tested
strengths to the predictions of ACI 349,8 Eurocode 2,12
Eurocode 8,11 and Harries et al.17 (Eq. (A1), (A3), (A4), and
(A5)). The test results included results of a previous study.18
In most existing test specimens,6,15,21,23,24 sliding failure
occurred after flexural yielding. Thus, only the test results by
Kono et al.,18 which showed sliding failure before flexural
yielding, were included. Specimen H0.5MG-C, showing
flexural yielding without sliding, was excluded. In the ACI
349 prediction, the maximum limit of yield strength, fyv ≤
420 MPa (60 ksi), was not used. Figures 9(a) to 9(d) show
that in the majority of the specimens, the peak strength did
not reach the design shear-friction strength due to the effects
of cyclic loading and high strength re-bars.
Fig. 8—Measured strains of vertical bars in specimens. The bar stress fsf at sliding failure can be estimated by
(Note: 1 mm = 0.0394 in.; 1 kN = 0.225 kip.) applying the maximum tested strength Vtest,s to the nominal
shear-friction equation by ACI 349, Vsf: fsf = (Vtest,s – μP)/
In Specimens H0.5MU (Fig. 8(b)), H0.5HU, H0.5QU,
(μAvall), where Avall is sum of bar area of vertical bars and
H0.5MU-C, H0.33MU (Fig. 8(c)), and H0.33HU, with
additional shear-friction bars (that is, Avall = Av + Asf), and
the shear-sliding failure mode, the strain of the horizontal
P is axial force (positive in compression). The maximum
bars was smaller when the displacement at sliding failure
tested strength is defined as the smaller of the two maximum
was greater. This result indicates that the shear force was
strengths in the positive and negative directions, Vtest,s
limited by the sliding failure at the wall-foundation inter-
(Table 4). In the evaluation, Avall were assumed to contribute
face. In H0.33MU-AS, H0.33HU-AS, H0.33MU-AL
to the shear-friction strength. The bar stresses fsf were esti-
(Fig. 8(d)), and N0.33MU-AL, with additional shear-friction
mated as 339 to 655 MPa (49.2 to 95.0 ksi) (Table 4), except
bars, the strain of the horizontal bars was greater than that of
H0.5MG-C with flexural yielding. In most of the specimens,
the specimens without additional shear-friction bars.
the estimated bar strengths did not reach the yield strength
of 550 MPa (80 ksi).
Figure 10 shows the relationships between the test
strengths Vtest,s and the area of overall vertical bars Avall (in

74 ACI Structural Journal/January 2018


Fig. 9—Comparison of test shear strengths and predictions of current design codes. (Note: 1 mm2 = 0.0016 in.2)
3. Regardless of the wall aspect ratios, the maximum
tested strength was proportional to the area of overall
vertical bars. This result indicates that the shear-friction
strength was not directly related to the magnitude of flex-
ural moment, as pointed out by Mattock et al.,13 because
flexural moment induces both compression and tension in a
cross-section, which have a positive and negative effect on
the shear-friction resistance, respectively. Thus, the effect of
flexural moment can be neglected.
From the trend line and the test results of Characteristics
1 to 3 listed previously, the shear-friction strength can be
defined as follows. For untreated interface (μ = 0.6)

Vsf2 = 250Avall = μ(420)Avall (μ = 0.6) (SI)


(1a)
Vsf2 = 36Avall = μ(60)Avall (μ = 0.6) (U.S. customary)

Fig. 10—Evaluation of design strength of shear-friction For yield strengths less than 420 MPa (60 ksi), clearly the
bars. (Note: 1 mm2 = 0.0016 in.2; 1 kN = 0.225 kip; 1 MPa bar stress cannot be greater than the yield strength. Further,
= 0.145 ksi.) the effect of compression force, if any, should be considered.
Thus, the shear-friction strength can be redefined as follows
the case of compression, the effective area = [Avallfyv + P]/
fyv = Avall + P/fyv). In Fig. 10, regardless of the steel grade of
Vsf2 = μ(Avallfyv + P) (fyv ≤ 420 MPa [60 ksi]) (1b)
the bars, the maximum tested strength was approximately
proportional to the bar area. In the case of the axial compres-
Equation (1b) is identical to the current ACI 349 equation.
sion (0.07Acfc′, H0.5MU-C) and surface treatment (μ = 1.0,
Particularly, the yield strength of vertical bars is limited to
H0.33MR), the maximum strength was greater. Exception-
420 MPa (60 ksi), as defined in the shear-friction design of
ally in the case of H0.33MU-AL with HD10 additional
ACI 349. A previous study4 reported that Grade 550 MPa
shear-friction bars of 300 mm (11.82 in.) anchorage length,
(80 ksi) bars are applicable to the flexural strength and shear
the maximum tested strength was significantly greater than
strength of walls. However, the test result in this study indi-
the proportional trend line.
cates that the design yield strength of bars for shear-friction
From the test results of the specimens with untreated
design should be limited to 420 MPa (60 ksi), although
interfaces (Fig. 10), the characteristics of the shear-friction
550 MPa (80 ksi) bars are used for flexural and shear design.
strength under cyclic loading can be summarized as follows:
Figure 9(e) shows the strength ratios of the test specimens
1. In the case of Grade 420 and 550 MPa (60 and 80 ksi)
predicted by Eq. (1), including the test results by Kono et
bars, the shear-friction strength was proportional to the area
al.18 In Fig. 9(e), the shear-friction strength by Eq. (1) safely
of overall vertical bars, showing a trend line with the slope of
predicted the test strength. The exception was H0.5HU and
250 MPa (36.3 ksi), regardless of the yield strength of bars.
H0.33HU, where the strength ratio Vtest,s/Vsf2 was 0.88 and
2. The trend line passes through the origin, which indi-
0.81, respectively. In the specimen, the maximum loads in
cates that the cohesion resistance was negligible due to the
the positive and negative directions significantly differed,
effect of cyclic loading.

ACI Structural Journal/January 2018 75


which indicates that an unexpected loading condition such to 420 MPa (60 ksi), as specified in ACI 349 (or ACI 318);
as torsion occurred under the pull-push lateral loading. 2) the effect of flexural moment can be neglected, as pointed
out by Mattock et al.13; and 3) due to the effect of cyclic
CONCLUSIONS loading, the cohesion resistance is negligible.
To investigate the shear-friction strength of walls with 7. In low-rise walls with construction joints, the safety
Grade 550 MPa (80 ksi) bars, five wall specimens with hw/lw margin of shear-friction strength is smaller than that of
= 0.5 and eight wall specimens with hw/lw = 0.33 were tested flexural strength. Further, the use of additional vertical bars
under cyclic lateral loading. In most specimens, surface increases the flexural strength as well as the shear-fric-
treatment in the interface of the construction joint was not tion strength. Thus, the failure mode of shear walls with
used: μ = 0.6 (ACI 349). The major findings of the present construction joints is expected to be determined by shear-
study are summarized as follows: sliding before flexural yielding.
1. Sliding failure occurred in all specimens with a construc-
tion joint at the wall-foundation interface. In the speci- AUTHOR BIOS
mens with additional shear-friction bars (H0.33MU-AS, Jang-Woon Baek is a Researcher at the Institute of Engineering Research
at Seoul National University, Seoul, South Korea. He received his BE, MS,
H0.33HU-AS, and N0.33MU-AL), a new horizontal crack and PhD from the Department of Architecture & Architectural Engineering
was formed along the end of the additional bars and sliding at Seoul National University.
failure occurred along the horizontal crack. This result indi-
ACI member Hong-Gun Park is a Professor in the Department of Architec-
cates that the major failure mode of low-rise walls with ture & Architectural Engineering at Seoul National University. He received
construction joints is shear sliding failure rather than flexural his BE and MS in architectural engineering from Seoul National University
yielding mode. and his PhD in civil engineering from the University of Texas at Austin,
Austin, TX. His research interests include inelastic analysis and the seismic
2. On the other hand, H0.5MG-C, which had a groove design of reinforced concrete structures.
construction joint at the interface, showed a typical flex-
ure-shear failure mode without shear sliding. Byung-Soo Lee is a Researcher in the Plant Construction & Engineering
Laboratory at Central Research Institute of Korea Hydro & Nuclear Power
3. In most of the specimens, the maximum tested strength Co., Ltd., Gueongju, South Korea. He received his BE in architectural engi-
was proportional to the vertical reinforcing bar ratios, neering from Kyung Hee University, Seoul, South Korea.
regardless of the reinforcement grade and wall aspect ratio
Hyun-Mock Shin is a Professor in the Department of Civil and Environ-
(hw/lw = 0.33 or 0.5). The effect of axial compression on mental Engineering at Sungkyunkwan University, Seoul, South Korea. He
the shear-friction strength agrees with the prediction of the received his BE and MS in civil engineering from Seoul National Univer-
ACI 349 shear-friction equation. By roughening the surface sity and his PhD in civil engineering from the University of Tokyo, Tokyo,
Japan. His research interests include nonlinear structural analysis and the
at wall-foundation interface complying with the “rough- seismic design of reinforced concrete structures.
ened surface” specified by ACI 349 and Eurocode 2, the
maximum strength was increased approximately as twice as ACKNOWLEDGMENTS
that of the specimen with a “smooth surface.” This work was supported by Korea Hydro & Nuclear Power Co. Ltd
4. In the specimens with Grade 550 MPa (80 ksi) bars (No. L16S140000) and by the Nuclear Power Core Technology Develop-
ment Program of the Korea Institute of Energy Technology Evaluation and
and sliding failure mode, the maximum tested strength Planning (KETEP), granted financial resource from the Ministry of Trade,
was smaller than the nominal shear-friction prediction by Industry & Energy, Republic of Korea (No. 2014151010169B). The Insti-
ACI 349 (Vtest/Vsf = 0.65 to 0.83; yielding of Grade 550 MPa tute of Engineering Research at Seoul National University also provided
research facilities for this work.
[80 ksi] bars was assumed to be valid). On the other hand,
in the specimens with Grade 420 MPa (60 ksi) bars, the
REFERENCES
maximum tested strength was close to or greater than the 1. Cheng, M.-Y.; Hung, S.-C.; Lequesne, R. D.; and Lepage, A., “Earth-
nominal shear-friction prediction by ACI 349 (Vtest/Vsf = 0.99 quake-Resistant Squat Walls Reinforced with High-Strength Steel,” ACI
to 1.06). Structural Journal, V. 113, No. 5, Sept.-Oct. 2016, pp. 1065-1076. doi:
10.14359/51688825
5. In the specimens with additional shear-friction bars of 2. Kabeyasawa, T., and Hiraishi, H., “Tests and Analyses of High-
a long anchorage length, the strength ratios were greater Strength Reinforced Concrete Shear Walls in Japan,” High-Strength
than those using the short anchorage length. The test result Concrete in Seismic Regions, SP-176, C. French and M. Kreger, eds., Amer-
ican Concrete Institute, Farmington Hills, MI, 1998, pp. 281-310.
indicates that the development length of the shear friction 3. Maier, J., and Thürlimann, B., “Bruchversuche an Stahlbetonscheiben
reinforcing bars should be greater than the ordinary devel- (in German),” Institut für Baustatik und Konstruktion, Eidgenössische
opment length specified by current design codes, because Technische Hochschule (ETH) Zürich, Zürich, Switzerland, 1985, 130 pp.
4. Park, H. G.; Baek, J. W.; Lee, J. H.; and Shin, H. M., “Cyclic Loading
in walls subjected to cyclic loading, wide and severe flex- Tests for Shear Strength of Low-Rise Reinforced Concrete Walls with
ure-shear cracks degrade the bond strength of the shear Grade 550 MPa Bars,” ACI Structural Journal, V. 112, No. 3, May-June
friction reinforcing bars. In the present study, development 2015, pp. 299-310. doi: 10.14359/51687406
5. Saito, H.; Kikuchi, R.; Kanechika, M.; and Okamoto, K., “Experi-
length of 30-bar diameter is recommended for the U-type mental Study on the Effect of Concrete Strength on Shear Wall Behavior,”
Grade 550 MPa (80 ksi) shear-friction reinforcement. Proceedings of the Tenth International Conference on Structural Mechanics
6. The shear-friction strengths of the test specimens with in Reactor Technology, V. H08/05, Anaheim, CA, 1989, pp. 227-232.
6. Paulay, T.; Priestley, M.; and Synge, A., “Ductility in Earthquake
a construction joint were proportional to μAvallfs (μ = 0.6, Resisting Squat Shearwalls,” ACI Journal Proceedings, V. 79, No. 4, Apr.
fs = 420 MPa [60 ksi]), regardless of the yield strength of 1982, pp. 257-269.
bars and wall aspect ratio. The result indicates that: 1) the 7. Wasiewicz, Z. F., “Sliding Shear in Low Rise Shear Walls under
Lateral Load Reversals,” master’s thesis, University of Ottawa, Ottawa,
design yield stresses of shear-friction bars should be limited ON, Canada, 1988, 127 pp.

76 ACI Structural Journal/January 2018


8. ACI Committee 349, “Code Requirements for Nuclear Safety-Related d) Concrete anchored to as-rolled structural steel by
Concrete Structures (ACI 349-13) and Commentary,” American Concrete
Institute, Farmington Hills, MI, 2013, 196 pp.
headed studs or by reinforcing bars: μ = 0.7.
9. ACI Committee 318, “Building Code Requirements for Structural In ACI 318, the permissible maximum shear-friction strength
Concrete (ACI 318-14) and Commentary (ACI 318R-14),” American is specified as follows: for normalweight concrete either placed
Concrete Institute, Farmington Hills, MI, 2014, 520 pp.
10. Birkeland, P. W., and Birkeland, H. W., “Connections in Precast
monolithically or placed against hardened concrete with clean
Concrete Construction,” ACI Journal Proceedings, V. 63, No. 3, Mar. 1966, surface free of laitance and intentionally roughened to a full
pp. 345-368. amplitude of approximately 6 mm (0.25 in.)
11. British Standards Institution, “Eurocode 8: Design of Structures for
Earthquake Resistance,” European Committee for Standardization, Brus-
sels, Belgium, 2003, 229 pp. Vsf,max = min[0.2fc′Acv, (3.3 + 0.08fc′)Acv, 11Acv] (SI) (A2a)
12. British Standards Institution, “Eurocode 2: Design of Concrete
Structures – Part 1-1: General Rules and Rules for Building (EN-1992-1-
1:2004:E),” European Committee for Standardization, Brussels, Belgium,
For all other cases
2004, 225 pp.
13. Mattock, A. H.; Johal, L.; and Chow, H. C., “Shear Transfer in Rein- Vsf,max = min[0.2fc′Acv, 5.5Acv] (SI) (A2b)
forced Concrete with Moment or Tension Acting across the Shear Plane,”
PCI Journal, V. 20, No. 4, 1975, pp. 76-93.
14. Wood, S. L., “Shear Strength of Low-Rise Reinforced Concrete where fc′ is cylinder strength of concrete (MPa); and Acv is
Walls,” ACI Structural Journal, V. 87, No. 1, Jan.-Feb. 1990, pp. 99-107. total sectional area of the shear interface (mm2). The yield
15. Orakcal, K.; Massone, L. M.; and Wallace, J. W., “Shear Strength
of Lightly Reinforced Wall Piers and Spandrels,” ACI Structural Journal,
strength of shear-friction bars is limited to 420 MPa (60 ksi)
V. 106, No. 4, July-Aug. 2009, pp. 455-465. in both ACI 3189 and ACI 349.8
16. Bass, R. A.; Carrasquillo, R. L.; and Jirsa, J. O., “Shear Transfer In ACI 349, for conservative design of NPP walls, the
across New and Existing Concrete Interfaces,” ACI Structural Journal,
V. 86, No. 4, July-Aug. 1989, pp. 384-394.
permissible maximum shear-friction strength is specified by
17. Harries, K. A.; Zeno, G.; and Shahrooz, B., “Toward an Improved using Eq. (A2b) for all cases.
Understanding of Shear-Friction Behavior,” ACI Structural Journal,
V. 109, No. 6, Nov.-Dec. 2012, pp. 835-844.
18. Kono, S.; Tanaka, H.; and Watanabe, F., “Interface Shear Transfer for
Eurocode 2
High Strength Concrete and High Strength Shear Friction Reinforcement,” In Eurocode 2,12 the shear strength at an interface between
Proceedings of High Performance Materials in Bridges, 2003, pp. 319-328. concrete cast at different times is defined as the sum of
19. Vintzeleou, E., and Tassios, T., “Behavior of Dowels under Cyclic Defor-
mations,” ACI Structural Journal, V. 84, No. 1, Jan.-Feb. 1987, pp. 18-30.
concrete cohesion and friction, as follows
20. Figueira, D.; Sousa, C.; Calçada, R.; and Neves, A. S., “Push-off
Tests in the Study of Cyclic Behavior of Interfaces between Concretes Cast VEC2 = cAcv fct + μf(Avf fyv + P) ≤ 0.5vfc′ (SI) (A3)
at Different Times,” Journal of Structural Engineering, ASCE, V. 142,
No. 1, 2016, 04015101. doi: 10.1061/(ASCE)ST.1943-541X.0001364
21. Salonikios, T. N.; Kappos, A. J.; Tegos, I. A.; and Penelis, G. G., where fct is tensile strength of concrete (MPa); v is strength
“Cyclic Load Behavior of Low-slenderness Reinforced Concrete Walls: reduction factor; and c and μf are factors that depend on the
Design Basis and Test Results,” ACI Structural Journal, V. 96, No. 4,
July-Aug. 1999, pp. 649-661.
roughness of the interface as follows.
22. Hawkins, N. M., and Ghosh, S. K., “Acceptance Criteria for Special a) Indented: A surface with indentations complying with
Precast Concrete Structural Walls Based on Validation Testing,” PCI Fig. 6.9 in Eurocode 2: c = 0.5 and μf = 0.9
Journal, V. 49, No. 5, 2004, pp. 78-92. doi: 10.15554/pcij.09012004.78.92
23. Greifenhagen, C., and Lestuzzi, P., “Static Cyclic Tests on Lightly
b) Rough: A surface with at least 3 mm (0.12 in.)
Reinforced Concrete Shear Walls,” Engineering Structures, V. 27, No. 11, roughness at about 40 mm (1.58 in.) spacing, achieved by
2005, pp. 1703-1712. doi: 10.1016/j.engstruct.2005.06.008 raking, exposing of aggregate or other methods giving an
24. Athanasopoulou, A., and Parra-Montesinos, G., “Experimental Study
on the Seismic Behavior of High-Performance Fiber-Reinforced Concrete
equivalent behavior: c = 0.40 and μf = 0.7
Low-Rise Walls,” ACI Structural Journal, V. 110, No. 5, Sept.-Oct. 2013, c) Smooth: A slipformed or extruded surface, or a free
pp. 767-777. surface left without further treatment after vibration: c =
0.20 and μf = 0.6
APPENDIX A: SHEAR-FRICTION STRENGTHS OF d) Very smooth: A surface cast against steel, plastic, or
CURRENT DESIGN CODES specially prepared wooden molds: c = 0.025 to 0.10 and
ACI 349 and ACI 318 μf = 0.5.
In ACI 3498 (or ACI 3189), the shear-friction strength of a Generally, the shear-friction predictions by Eq. (A3) are
construction interface is defined as follows greater than those by Eq. (A1) because of the inclusion of the
concrete cohesion term.
Vsf = μ(Avf fyv + P) (A1)
Eurocode 8
where Avf is area of shear-friction reinforcement across the In Eurocode 8,11 the shear resistance against sliding is
shear plane (mm2); fyv is yield strength of the shear-friction defined as the sum of the dowel resistance of vertical bars
reinforcement (MPa); P is applied axial compression force (Vdd), shear resistance of inclined bars (Vid) (at an angle φ to
(N); and μ is friction coefficient defined as follows. the potential sliding plane—for example, construction joint),
a) Concrete placed monolithically: μ = 1.4 and friction resistance (Vfd)
b) Concrete placed against hardened concrete with surface
intentionally roughened: μ = 1.0 VEC8 = Vdd + Vid + Vfd (A4a)
c) Concrete placed against hardened concrete not
intentionally roughened: μ = 0.6 (
Vdd = min 1.3∑ Asj f c′ f yv , 0.25 f yv ∑ Asj ) (SI) (A4b)

ACI Structural Journal/January 2018 77


Vid = ∑Asi fyvcosφ (A4c) where fc′ is cylinder compressive strength of concrete
(MPa); lw is overall length of the wall (mm); h is thickness
Vfd = min(μf[(∑Asj fyv + Ned)ξ + MEd/z], 0.5η fc′ξlwbw) (SI) of the wall (mm); d is distance from the extreme compression
(A4d) fiber to the centroid of longitudinal tension reinforcement
where ∑Asj is the sum of the areas of the vertical bars of (mm) (= 0.8lw in ACI 349); Nu is axial force (positive sign in
the web and of additional bars arranged in the boundary compression) [N]; Mu is applied flexural moment (N·mm); Vu
elements specifically for resistance against sliding (mm2); is applied shear force (N); As is area of shear reinforcement
∑Asj is the sum of the areas of all inclined bars in both (mm2) within spacing s (mm); and fyh is specified yield strength
directions (mm2); φ is an angle to the potential sliding plane of web shear reinforcement (MPa). Vn in Eq. (B1) shall not
(rad); μf is concrete-to-concrete friction coefficient under exceed the permissible maximum shear strength 0.83√fc′hd
cyclic actions, which may be assumed equal to 0.6 for (Vn,max). In ACI 349 (or ACI 318), the concrete shear strength
smooth interfaces and to 0.7 for rough interfaces, as defined Vc of walls is determined as the smaller of the values of
in Eurocode 2; Ned and Med are the design axial load (N) Eq. (B2b) and Eq. (B2c) unless Eq. (B2a) is used. Eq. (B2a)
and design flexural moment (N·mm), respectively, obtained is generally regarded as the basic concrete shear strength of
from the analysis for seismic design; ξ is normalized neutral flexural members including squat walls. Thus, in the present
axis depth; z is length of the internal lever arm (mm); η = study, for simplicity, Eq. (B2a) was used for the concrete shear
0.6(1 – fc′/250); lw is length of wall (mm); and bw is depth of strength. In this case, the permissible maximum shear strength
wall (mm). In Eurocode 2 and Eurocode 8, the coefficients of horizontal web reinforcing bars is defined as follows
of c and μ for concrete monolithically cast are not defined.
Vsmax = 0.83 f c′hd − Vc = 0.66 f c′hd (SI) (B4)

Harries’s equation
Harries et al.17 studied the effect of high-strength shear- On the other hand, in the seismic provision of ACI 349
friction bars, and proposed a shear-friction strength as follows (ACI 318), the shear strength of walls is specified as follows

VHarries = cAcv fc′ + 0.002AvfEs ≤ 0.20Acvfc′ (SI) (A5) Vseis = Acv (α c f c′ + ρh f yh ) (B5)

where α = 0.075, 0.040, and 0 for the interface with The first term on the right-hand side indicates Vc, and
monolithically cast uncracked, cold-jointed, and monolithically the second term indicates Vs. In Eq. (B5), Acv is the total
cast precracked conditions, respectively; and Es is elastic sectional area, and the coefficient αc is 0.25 (3.0 in U.S.
modulus of vertical bars (MPa). In Eq. (A5), the shear-friction customary units) for hw/lw ≤ 1.5.
strength is determined by the elastic modulus Es of the shear- In the seismic provision, the permissible maximum shear
friction bars, regardless of the yield strength. strength is specified as follows

APPENDIX B: MAXIMUM SHEAR Vseis,max = 0.66 f c′Acv (SI) (B6a)



REINFORCEMENT RATIO
In NPP walls, because of the high safety requirements, In the case of walls with a rectangular cross-section, Acv =
the shear reinforcement ratio for the design is close to the hlw and d = 0.8lw (ACI 349 or ACI 318). Thus, the permissible
maximum shear reinforcement ratio specified by current maximum shear strength of walls with a rectangular cross
design codes. Thus, most of the specimens tested in the section can be refined as follows
present study were designed with the permissible maximum
shear reinforcement ratio. Vseis,max = 0.66 f c′Acv = 0.66 f c′hlw = 0.825 f c′hd (SI)
In ACI 3498 (or ACI 3189), the shear strength of walls is (B6b)
defined as the sum of the contributions of concrete Vc and
shear reinforcement Vs From Eq. (B5) and (B6), the permissible maximum shear
strength of horizontal web reinforcing bars for hw/lw ≤ 1.5 is
Vn = Vc + Vs (SI) (B1) defined as follows

Vc = 0.17 f c′hd (SI) (B2a) Vsmax = (0.66 − 0.25) f c′Acv = 0.5125 f c′hd (SI) (B7)

Comparing Eq. (B4) and Eq. (B7), the greater value


Vc = 0.27 f c′hd + N u d /4lw (SI) (B2b)
of Vsmax in Eq. (B4) was used for the design of the test
specimens. Substituting Vsmax in Eq. (B4) into Vs in Eq. (B3),
 l (0.1 f c′ + 0.2 N u /(lw h))  the permissible maximum horizontal reinforcing bar ratio ρh
Vc = 0.05 f c′ + w  hd (SI) (B2c)
 ( M u /Vu − lw /2)  was calculated. The vertical web reinforcing bar ratio ρv was
determined to be the same as ρh according to the seismic
provisions of ACI 349 (or ACI 318): ρv ≥ ρh unless the aspect
Vs = As fyhd/s (B3)
ratio of walls is greater than 2.0.

78 ACI Structural Journal/January 2018


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 115-S06

Behavior of Straight and T-Headed ASTM A1035/A1035M


Bar Splices in Flexural Members
by Sergio F. Breña, Jeffrey Messier, and Sean W. Peterfreund

Interest has increased in recent years in the use of high-strength several documents have been developed that identify key
reinforcement in structural concrete members. This paper discusses issues that need to be considered and those that need further
the behavior of flexural elements that contain coupled or lapped study (ACI Innovation Task Group 6 2010; Applied Tech-
high-strength reinforcement conforming to ASTM A1035/A1035M. nology Council 2014). For flexural applications, the ACI
Behavior of elements containing couplers or spliced straight and
ITG-6R-10 design guide (ACI Innovation Task Group 6 2010)
T-headed bars are included, adding to the number of test results
provides key information that should be considered for design
available for broadening the application of high-strength reinforce-
ment in structural concrete. In addition to general observations on of beams containing ASTM A1035/A1035M reinforcement.
the behavior of specimens tested in this experimental program, Pertaining to straight development length of ASTM A1035/
information on measured strain distributions along the splices A1035M bars in tension, this document recommends using
are provided from which key force-transfer mechanisms between Eq. (4-11a) in ACI 408R-03 (Joint ACI-ASCE Committee
bars are identified. The benefits of having transverse reinforcement 408 2003) (rewritten as follows with factors for bar location,
within the splice of T-headed are highlighted. epoxy coating, and lightweight concrete set equal to 1.0) with
a strength reduction factor ϕ equal to 0.80 instead of 0.82
Keywords: high-strength reinforcement; lap splices; T-headed reinforcement.

 fy 
INTRODUCTION
ld  f ′1/ 4 − φ2400ω
Interest has increased in recent years in the use of high- = c
(1)
strength reinforcement in structural concrete to facilitate db  cω + K tr 
construction operations and decrease congestion in heavily φ76.3  
 d b
reinforced concrete members (ACI Innovation Task Group 6
2010). Limitations on the use of high-strength reinforcing The yield strength fy used in this equation should be
bars conforming to ASTM A1035/A1035M (Grades 100 established following the 0.2% offset method (refer to the
and 120) are placed in the ACI 318-14 building code (ACI Notation section for definition of variables). Development
Committee 318 2014), both in type of application and lengths obtained using this equation, however, can be signif-
usable strength. ASTM A1035/A1035M bars do not exhibit icant, particularly for large-diameter bars.
a well-defined yield plateau so yield strength is commonly T-headed reinforcement may be used to decrease the length
determined using the 0.2% offset method. These bars typi- required to develop high-strength reinforcement, thereby
cally have a 0.2% offset yield strength that exceeds 100 ksi addressing some of the concerns associated with the long
(690 MPa). However, ACI 318-14 places limits on the development lengths of straight bars. Current provisions
stress that can be used when designing with high-strength in ACI 318-14 Section 25.4.4.2 are limited to reinforce-
reinforcement because of limited availability of test data. ment with a nominal yield strength not exceeding 60 ksi
For example, ACI 318-14 sets an upper value to the yield (420 MPa) because the experiments upon which the provi-
strength of spiral reinforcement for confinement to 100 ksi sions are based were limited to this reinforcement grade.
(690 MPa), and lower stress limits are imposed when using Furthermore, a clear spacing between adjacent bars less
high-strength reinforcement in other applications. than 4db is currently not allowed in ACI 318-14 because
The use of high-strength reinforcement in flexural appli- of concerns that a change in the force-transfer mechanisms
cations has been limited to date primarily because of lack of between spliced bars might affect splice strength. This
design guidance for its use. Given the higher yield stress of spacing limit prevents T-heads from contacting adjacent bars
high-strength reinforcement, cracks forming at service load and requires a wider space to accommodate splices. Exper-
levels could be wider than those that form when normal- imental data are needed to broaden current design provi-
strength steel is used, raising concerns about ingress of water sions for spliced T-headed reinforcement to applications that
and corrosive agents into the concrete element. Furthermore, involve high-strength reinforcement.
the longer embedment required to develop higher tensile
stresses might cause reinforcement congestion and generate
designs that become impractical. Therefore, methods to that ACI Structural Journal, V. 115, No. 1, January 2018.
allow high-strength reinforcement to develop its strength MS No. S-2016-298.R1, doi: 10.14359/51700782, received February 20, 2017, and
reviewed under Institute publication policies. Copyright © 2018, American Concrete
without creating congestion are needed. Institute. All rights reserved, including the making of copies unless permission is
In response to the need to increase the use of high-strength obtained from the copyright proprietors. Pertinent discussion including author’s
closure, if any, will be published ten months from this journal’s date if the discussion
reinforcement in structural concrete design applications, is received within four months of the paper’s print publication.

ACI Structural Journal/January 2018 79


Fig. 1—Geometry and reinforcement of Group L specimens. Fig. 2—Geometry and reinforcement of Groups C and T
specimens.
RESEARCH SIGNIFICANCE
This research provides needed data on the behavior of of 1.5 in. (38 mm) around all reinforcing bars was provided
flexural elements that contain ASTM A1035/A1035M bars in all specimens.
as the main tensile reinforcement. Laboratory experiments Specimens in Group L were longer than specimens in
focused on investigating the splice behavior of bars with Groups C and T in order to fit the required lap splice length
couplers, straight lap splices, and splices between T-headed within the maximum moment region of the beams. Splicing
high-strength bars. The findings from this research will bars with mechanical couplers or T-heads allowed short-
complement existing and future investigations that may be ening of the maximum moment region of the beams and,
used to develop design provisions to increase use of high- therefore, the overall length of these specimens was shorter
strength bars in a wider variety of applications than those (Groups C and T).
currently permitted in building and bridge codes. A four-point testing configuration was used to generate
a constant moment region in the center of the beams. The
EXPERIMENTAL PROGRAM beams were tested in an inverted position (tension on top)
Specimen description to facilitate observation and mapping of the progression of
Laboratory tests in this research project included 15 cracking throughout the tests. Beams were supported using
beams containing ASTM A1035/A1035M longitudinal rein- rollers placed at the end of the constant moment region, and
forcement on the tension side of the beams spliced in the loading was applied in the downward direction near the end
region of maximum moment. The specimens were organized of the beams using hydraulic jacks. Bar splices were located
in three groups, depending on splice type. Six specimens within the center region of the beams (refer to Fig. 1 and 2).
had straight-bar lap splices, each with a different overlap- Specimen geometry and reinforcement are shown in Fig. 1
ping length (Group L). Reinforcement in eight of the nine for Group L specimens and Fig. 2 for Groups C and T speci-
remaining specimens was spliced using threaded couplers or mens. Specimen dimensions are summarized in Table 1. The
T-heads (Groups C and T, respectively). One specimen was span length of Group L specimens was 168 in. (4200 mm),
built with continuous tension bars as a control for specimens which included a length of the constant moment region equal
using No. 5 bars in Groups C and T. to 68 in. (1700 mm). The span length in Group T and C spec-
Specimen notation consists of three groups of alphanumeric imens was 114 in. (2900 mm) with a length of the constant
characters separated by dashes. The first group denotes the moment region equal to 30 in. (760 mm). Cross-sectional
specimen group (L, C, or T), the second group of characters dimensions of Group L specimens were 13 in. wide by
denotes the bar size (No. 4, 5, or 6), and the last character 9 in. deep (330 x 229 mm); those of Group C specimens were
group denotes the coupler manufacturer (M or E), or the length either 11 or 9 in. wide (279 or 229 mm) by 14 in. (356 mm)
of the splice (in.) and presence of transverse reinforcement deep; and those of Group T specimens were 9 in. wide by
along the splice (R). For example, Specimen T-5-12 denotes 14 in. deep (229 x 356 mm). The small width variation in
a specimen in Group T (T-headed splice) with two No. 5 bars Group C specimens was to provide the same lateral spacing
as main tension reinforcement and a 12 in. (305 mm) splice. between bars as a function of bar diameter for the two different
Geometry and reinforcement of the specimens is illus- bar sizes tested (No. 6 [19 mm] and No. 5 [16 mm]).
trated in Fig. 1. The non-contact spliced bars in Group L Main tensile reinforcement in Group L specimens consisted
specimens had a transverse clear spacing of one bar diameter of two No. 4 (13 mm) or two No. 5 (16 mm) bars. The main
db and a minimum of 2db between the groups of longitudinal reinforcement was lap spliced within the constant moment
bars being spliced. Splices in specimens with T-headed rein- region using different lengths in each specimen. For each
forcement were constructed so that the head of the tee was bar size (No. 4 [13 mm] or No. 5 [16 mm]), the longest lap
in contact with the adjacent bar being spliced. A clear cover splice length chosen was consistent with ACI 318-14 devel-
opment length equations in Section 25.4.2.3, including the

80 ACI Structural Journal/January 2018


Table 1—Properties of laboratory specimens
Specimen b, in. h, in. As, in.2 ls, in. ld,* in. fc′, psi fy, ksi fu, ksi εy εu
L-4-24 13 9 0.4 24 6190 0.007 0.065
L-4-19 13 9 0.4 19 18.3 6210 120 198 0.007 0.065
L-4-14 13 9 0.4 14 6085 0.007 0.065
L-5-37 13 9 0.62 37 6190 0.007 0.060
L-5-29 13 9 0.62 29 23.5 6155 120 178 0.007 0.060
L-5-21 13 9 0.62 21 6830 0.007 0.060
Control CT 9 14 0.62 — — 5460 105 155 0.0056 0.060
C-5-M 9 14 0.62 — — 5140 105 155 0.0056 0.060
C-6-M 11 14 0.88 — — 5400 121 170 0.0062 0.050
C-5-E 9 14 0.62 — — 4280
T-5-12 9 14 0.62 12 4320
T-5-8 9 14 0.62 8 4410
105 155 0.0056 0.060
T-5-4 9 14 0.62 4 15.8 4640
T-5-8R 9 14 0.62 8 4490
T-5-4R 9 14 0.62 4 4280
*
Computed using ACI 318-14 Eq. (25.4.2.3a) for L-series specimens and Section 25.4.4.2(a) for T-series specimens.
Notes: 1 in. = 25.4 mm; 145 psi = 1 MPa; 1 ksi = 1000 psi.

modification for a Class B lap splice (ACI 318-14, Section


25.5.2.1) to theoretically reach a bar stress approximately
equal to yield (120 ksi [827 MPa]); refer to “Material prop-
erties” section. To promote splice failure in other specimens,
splice lengths were reduced to approximately 80% and 60%
of the longest splice for each bar size.
Tensile reinforcement of Group T specimens consisted of
two No. 5 (16 mm) bars. The reinforcement in these speci-
mens was spliced using non-contact lapped bars that termi-
nated with a threaded T-head. The splice lengths in these
specimens were shorter than estimated using ACI 318-14
Section 25.4.4.2 to investigate the behavior of specimens up
until the onset of splitting failure. The longest splice length,
corresponding approximately to 75% of the code deter-
mined value, was equal to 12 in. (305 mm); this length was
decreased to 8 and 4 in. (204 and 102 mm) in other speci-
mens within this group. Previous researchers (Thompson et
al. 2006) have demonstrated that the presence of the T-head
allows bars to develop stress in a much shorter distance
compared with a straight bar splice. The T-headed ends of
adjacent spliced bars were staggered in the splice region, as
illustrated in Fig. 2.
Group C specimens had two No. 5 (16 mm) or two No. 6
(19 mm) bars as main tensile reinforcement. The bars were
spliced using threaded couplers machined from the same
material as the reinforcement. Longitudinal reinforcement
on the compression side in all specimens consisted of two
No. 3 (9 mm) bars. Transverse reinforcement (stirrups) Fig. 3—Splice details for Group T Specimens: (a) transverse
outside the constant moment region was fabricated using reinforcement (Specimens T-5-4R and 8R); and (b) strain
bars at spacing necessary to reach flexural strength without gauge placement.
prior occurrence of shear failure. stirrup to be placed at the center of the splice and the first
Only two specimens in Group T included transverse stirrup next to the T-head outside the splice. All other spec-
reinforcement within the lap splice (Fig. 3(a)). Specimens imens in Groups L and T had no transverse reinforcement
T-5-8R and T-5-4R had No. 3 stirrups at a spacing of 8 and within the lap splice or inside the constant moment region.
4 in. (200 and 100 mm), respectively, which allowed one

ACI Structural Journal/January 2018 81


Material properties
Beams were designed using a nominal 28-day compres-
sive strength of concrete, fc′, equal to 5000 psi (35 MPa) for
specimens in Group L and 4000 psi (28 MPa) for specimens
in Groups C and T. The tensile and compressive strengths of
concrete were determined the day specimens were loaded to
failure by testing companion cylinders and plain concrete
beams that were cast and cured under the same laboratory
conditions as the specimens. The concrete compressive
strength at the day of testing varied between an average
minimum of 4300 psi (30 MPa) and a maximum of 5460 psi
(38 MPa), depending on the time that elapsed between casting
and testing date. Average results from concrete cylinder and
beam tests for each specimen are listed in Table 1.
Longitudinal tension reinforcement conformed to ASTM
A1035/A1035M. Three different bar sizes were used in the
testing program, and each bar size had slightly different
material stress-strain properties. The stress-strain prop-
erties of two of the bar sizes (No. 5 [16 mm] and No. 6
[19 mm]) were determined in the laboratory by tensile testing
of bar segments cut from the same longitudinal reinforce-
ment used in the specimens. The stress-strain properties of the
No. 4 (13 mm) bars listed in Table 1 were reported by the
manufacturer of the reinforcing steel. Compression steel and
transverse reinforcement (No. 3 [9.5 mm] bars) conformed to
ASTM A615 Grade 60 (Grade 420); mechanical properties
were not determined in the laboratory for these bars. Measured
stress-strain curves of No. 5 (16 mm) and No. 6 (19 mm) bars
are shown in Fig. 4(a). The yield stress of the reinforcement
was determined using the 0.2% offset method as illustrated in
the figure. For use in moment-curvature analysis as described
later in this paper, relationships that approximately describe Fig. 4—Stress-strain properties of ASTM A1035/A1035M
the measured stress-strain behavior of the bars used in this bars: (a) measured; and (b) bilinear approximation. (Note:
testing program were obtained for the No. 5 (16 mm) bars 1 ksi = 6.895 MPa.)
constructed using Eq. (2a) and (2b) is presented in Fig. 4(b).
 29, 000ε s (ksi) ε s ≤ 0.00275 This comparison reveals that the bilinear approximation
 (2a)
fs =  0.4 would result in reasonably accurate stresses, particularly
163 − ε + 0.00194 (ksi) 0.00275 < ε s ≤ 0.060 within the post-yield region. The two stress-strain modeling
 s
approaches were used to calculate the maximum moment
that the sections could theoretically develop if the bars were
and for the No. 6 (19 mm) bars
continuous, as will be discussed later in this paper.
Commercially available couplers and T-heads are typi-
 29, 000ε s (ksi) ε s ≤ 0.00275
cally qualified to develop at least 1.25fy of reinforcement

fs =  0.4 (2b) with a well-defined yield plateau. Because of the high
178 − ε + 0.00132 (ksi) 0.00275 < ε s ≤ 0.060 strength characteristics of the ASTM A1035/A1035M bars
 s
used in this research, mechanical devices commercially
Equations (2a) and (2b) have the same form as those available for ASTM A615 bars could potentially fracture at
proposed by Mast et al. (2008) but with coefficients that low strains. Therefore, threaded mechanical couplers and
were adjusted to better represent the measured proper- T-heads were machined from round stock that had the same
ties of the bars used in this research. Stress-strain rela- material properties as the tension longitudinal reinforcement
tions proposed by Seliem et al. (2009) and Shahrooz et al. in this research. The geometry of couplers and T-heads fabri-
(2014) were also evaluated but were not selected because cated for this research project is presented in Fig. 5. Coupler
these material models gave a sharper transition from the E was only used in one of the specimens (Specimen C-5-E),
linear portion to the hardening region of the stress-strain as listed in Table 1. Ends of reinforcing bars were threaded at
diagrams. The stress-strain curves of the tension reinforcing the fabrication shop prior to delivery in accordance with the
bars were also modeled using a simple bilinear approxi- coupler or T-heads they were designed to use. The resulting
mation to simplify calculations. A comparison between the diameter of the couplers and T-heads was approximately
bilinear approximations for each bar size and the curves 50% larger than those used for ASTM A615 bars because
the higher material hardness generated difficulties during

82 ACI Structural Journal/January 2018


Fig. 5—Couplers and T-head geometries.
milling. Pictures illustrating coupled and T-head reinforced
bars during assembly of two selected specimens are shown
in Fig. 6.

Instrumentation
External instrumentation on the specimens consisted of
load cells and displacement transducers (linear potentiom-
eters). Individual load cells were used for each hydraulic
ram at the specimen ends to measure the total load applied.
Differences in measured load at each of the load cells were
less than 2% of the total applied load throughout testing,
so loading was considered symmetrically applied. Linear Fig. 6—Layout of coupled and T-headed reinforcing bar
potentiometers were positioned at beam centerline and at the cages showing strain gauges.
locations of load application near the beam ends. Given the
test configuration, the total beam deflection was determined of the T-head with the goal to separate the contribution of
by adding the average displacement measured near the ends the T-head from the stress developed along the spliced bars,
of the specimen to the displacement measured at centerline. but this was not always possible. Other strain gauges were
Bonded strain gauges were used to instrument reinforcing distributed at a constant spacing (between 2.5 and 3.0 in. [64
steel at various locations depending on each specimen and and 76 mm], if space was available) within the remaining
to instrument the compression surface of all specimens at length of the splice (Fig. 3).
midspan. Steel strain gauges were mounted on both rein-
forcing bars of Specimen CT at midspan. In Group C spec- TEST RESULTS
imens, steel strain gauges were bonded to both reinforcing A list of the failure modes observed after testing of each
bars 1.75 in. (44.5 mm) on either side from the center of specimen is given in Table 2. Control Specimen CT and
the coupler located at midspan. In specimens containing Group C (coupled bar) specimens, with exception of Spec-
splices (Group L and T), the spacing and number of steel imen C-5-M, exhibited flexural failure characterized by
strain gauges varied for each specimen because of differ- crushing of the concrete on the compression face of the
ences in splice length. Both bars in one of the two splices in beams after bar yielding. The bars in Specimen C-5-M frac-
Group L specimens were instrumented with equally spaced tured next to the threaded ends when the specimen reached
strain gauges. The spacing of gauges ranged from 2.75 in. 99% of the estimated flexural strength using the simpli-
(70 mm) for the specimen that had the shortest splice length fied approach described later in this paper. Concrete in the
(L-4-14) to 6.2 in. (157 mm) for the specimen containing compression zone had started to flake at this load level, indi-
the longest splice length (L-5-37). Although more closely cating impending crushing. Crack patterns at failure of all
spaced gauges are likely to better capture sudden changes in specimens containing splices were indicative of splitting
strain along the bars, it was decided to have the same number of concrete along the spliced bars. In specimens containing
of strain points along all spliced bars in this group. straight bar splices (Group L), flexural cracks formed first
In Group T specimens, three out of the four spliced within the constant moment region typically at or near
bars were instrumented using strain gauges at equal inter- midspan, and at the end of the splice. These cracks were
vals along the splice. The first strain gauge was placed at a followed by formation of longitudinal splitting cracks that
distance equal to 2db (1.25 in. [32 mm]) from the inside face appeared on the sides and/or on the tensile face of the beams
and gradually lengthened at higher loads. The splitting cracks

ACI Structural Journal/January 2018 83


Table 2—Summary of test results
Test results M-ϕ analysis Simplified
Ptest, Mtest, fs test, Mcalc, Mcalc, Pcalc, fs calc,†
Specimen kip kip-in. εs,test* ksi kip-in. Mtest/Mcalc kip-in. Mtest/Mcalc kip ksi fs test/fs calc Observed failure mode
L-4-24 14.2 370.7 0.0094 126 0.93 1.05 119 1.06 Bottom splitting
Bottom splitting + one
L-4-19 14.1 368.2 0.0107 132 0.92 1.04 106 1.25
399 353 14.1 side splitting
Bottom splitting + side
L-4-14 11.3 298.2 0.0073 118 0.75 0.84 91 1.30
splitting
Side and bottom splitting
L-5-37 19.6 505.7 0.0085 122 0.92 0.97 118 1.03
(corner)
548 523 20.9 Side and bottom splitting
L-5-29 18.9 488.2 0.0064 113 0.89 0.93 105 1.08
(corner)
L-5-21 14.5 378.2 0.0053 103 0.69 0.72 91 1.13 —
Flexural – concrete
Control CT 45.4 966.6 0.0122 135 928 1.04 941 1.03 44.8 141 0.96
crushing
C-5-M 44.9 957.6 0.0139 138 914 1.05 974 0.98 46.4 141 0.98 Reinforcing bar fracture
Flexural – concrete
C-6-M 58.1 1238.0 0.0127 135 1311 0.94 1355 0.91 64.5 147 0.92
crushing
Flexural –concrete
C-5-E 41.4 883.0 0.0168 143 1.01 0.96 140 1.02
crushing
T-5-12 20.2 439.1 0.0048 98 0.50 0.48 79 1.24 Side splitting along splice
T-5-8 14.6 320.5 0.0018 52 0.37 0.35 53 0.98 Side splitting along splice

871 916 43.6 Side and bottom splitting


T-5-4 13.7 301.5 0.0023 59 0.35 0.33 27 2.17
along splice
Side and bottom splitting
T-5-8R 23.2 501.4 0.0039 90 0.58 0.55 54 1.68
along splice
Bottom splitting
T-5-4R 21.1 457.8 0.0027 69 0.53 0.50 26 2.64
diagonally to T-head
*
εs,test is average of strains measured end of splice of all instrumented bars.

fs calc was determined using Eq. (3b) for Group L specimens, simplified flexural strength method for Group C specimens, and Eq. (4b) for Group T specimens.
Notes: 1 kip = 4.45 kN; 1 kip-in. = 113 kN-mm; 1 ksi = 6.895 MPa.

that formed during loading eventually led to the formation of Load-deflection response
a mechanism and failure of the splice by splitting. Measured peak loads and corresponding moments for all
Specimens in Group T also exhibited cracks that formed the specimens are summarized in Table 2. The moments
at the end of the splice region behind the T-headed end of listed in the table were adjusted to include the self-weight
alternate bars in addition to cracking at midspan. No longi- moment computed at the end of the constant moment region.
tudinal cracking along the reinforcing bars was apparent Load-deflection responses of various groups of specimens
throughout testing in these specimens. Failure of Group T are compared in Fig. 8. Deflection data were lost during
specimens was triggered by widening of the cracks at the testing of Specimen L-5-21, so the load-deflection curve
splice ends and sudden formation of longitudinal splitting for this specimen is not included in Fig. 8(a) but the peak
crack throughout the splice region. Inspection of the splice measured load is listed in Table 2. It should be noted that
region after failure of these specimens revealed the occur- the figures are drawn at different scales so that detailed
rence of a large displacement of the back side of the T-head features of the load-deflection plots of the different groups
as concrete crushed in front of the T-head (refer to Fig. 7). of specimens can be easily observed. The loads calculated
It is apparent that crushing of concrete released the T-head, under the assumption that specimens would reach their
causing a sudden increase in bond stresses along the bars, corresponding flexural strength at the critical section under
which led to the sudden splitting failure observed. The one of the load points Pcalc are indicated in each plot with a
formation of a truncated concrete cone between diagonal dashed horizontal line and are also listed in Table 2. These
cracks in front of the T-head was also visible in these spec- calculated loads were estimated by iterating on the neutral
imens (Fig. 7). The length of the cones measured in several axis position using the Whitney stress block for concrete in
specimens ranged between 2.1 and 2.2 in. (53 and 56 mm), compression and a bilinear approximation for the steel in
resulting in an angle of inclination between the bar axis and tension in accordance with Fig. 3(b). The measured material
the diagonal cracks of approximately 21 degrees. properties for different groups of specimens were used in
these calculations.

84 ACI Structural Journal/January 2018


Fig. 7—Failure of Specimen T-5-8R.

Fig. 8—Selected comparisons of load-deflection responses within groups of specimens.


For lap-spliced specimens, only Specimens L-4-24 and with fy = 120,000 psi (830 MPa) and the measured concrete
L-4-19 reached the estimated load Pcalc (Fig. 8(a)). These compressive strength is presented in Table 1. Only Speci-
specimens, however, failed by splitting along the lap splice mens L-4-14 and L-5-21 had ls/ld ratios less than 1.0, but it
almost immediately after reaching the estimated flexural should be noted that the splice factor of 1.3 was not used in
strength without a sizeable increase in deflection beyond the the calculation for ld.
peak load. Specimens L-5-37 and L-5-29 failed at approx- Specimens in Group C (coupled bar specimens) reached
imately 94% and 90% of the loads corresponding to the or exceeded the load corresponding to flexural strength
estimated flexural strength, respectively, while specimens calculated using measured material properties. Only Spec-
containing the shortest splice lengths (L-4-14 and L-5-21) imen C-6-M fell short of the calculated flexural strength, by
failed after only reaching approximately 80% and 70% approximately 10%.
of the calculated flexural strength, respectively. The ratio Specimens in Group T were deliberately designed to fail
of splice length ls of each specimen divided by calculated within the splice region prior to reaching the estimated flex-
development length ld using ACI 318-14 Section 25.4.2.3 ural strength of the cross section. Splice lengths were much

ACI Structural Journal/January 2018 85


Fig. 9—Strains in maximum moment region of Specimen CT and Group C specimens.
shorter than the tension development length for headed bars observed in all tests including Specimen C-5-M, which exhibited
computed using ACI 318-14 Section 25.4.4.2, assuming a reinforcing bar fracture. Plots showing strains measured on the
yield stress of 105 ksi (730 MPa) and the average concrete two bars on both sides of the coupler of all Group C specimens
strength of all Group T specimens. As shown in Fig. 8(c) are presented in Fig. 9(b) through Fig. 9(d). In general, all curves
and Table 2, all the specimens in Group T failed at a fraction lie close to each other for individual specimens with the excep-
of the capacity Pcalc calculated assuming continuous bars. tion of strains in one bar on one side of the coupler in Speci-
Figure 8(d) shows that specimens that contained transverse mens C-5-M and C-6-M. The differences in specific values could
reinforcement within the splice (T-5-8R and T-5-4R) were have been caused by the proximity of cracking to gauge loca-
able to develop between 55 and 60% higher ultimate loads, tions. However, the general trends in all specimens are similar
respectively, in comparison with corresponding specimens and compatible in appearance with strains measured at centerline
that did not contain transverse reinforcement along the splice of Control CT specimen. The similarity in observed strains indi-
(T-5-8 and T-5-4). Having transverse reinforcement within cated that couplers were able to effectively splice high-strength
the splice delayed failure and increased ductility, even with bars to reproduce the behavior of a beam with continuous bars.
the minimal amount of transverse reinforcement provided The strains developed along the splice in the two spec-
and the extremely short splices used in these specimens. imens that had the intermediate splice length (Specimens
L-4-19 and L-5-29) are shown in Fig. 10, and are repre-
Reinforcing bar and concrete strains sentative of other specimens in this group. Strains typically
Strains were measured within the maximum moment increased from the end of the bar to the end of the splice
region of all specimens. The location of strain gauges for region. Bar strains increased nonlinearly along the splice
each group of specimens was discussed in the “Instrumen- region, and local peaks were recorded at locations near
tation” section. concrete cracks. The curves are labelled a through e, corre-
Strains measured in Specimen Control CT and Group C sponding to the instrument location starting at the beginning
specimens are illustrated in Fig. 9. Failure associated of the splice as illustrated in the bottom right corner of each
with concrete crushing after reinforcing bar yielding was plot. The small drops in load at different points along the
observed in all these specimens as discussed previously, curves correspond to the load holds taken during testing to
so measured strains were anticipated to be compatible with mark cracks and document damage in the beams.
this failure mode. Concrete strains are shown as negative In Specimen L-4-19, the first change in slope occurred
values and were measured in the extreme compression face in all the strain curves at approximately 3 kip (13.3 kN)
at midspan; steel strains are positive and were measured at corresponds to first flexural cracking. Strains at location e
midspan for Specimen Control CT or next to the coupler increased markedly at approximately 7 kip (31 kN), corre-
for Group C specimens. As observed in the plots in Fig. 9, sponding to the occurrence of flexural cracking near the
steel strains exceeded 0.012 in tension and compressive strains instrument location. At loads nearing failure of the specimen
measured in the concrete reached the value of 0.003 often asso- above 10 kip (44.5 kN), strains increased at a higher rate
ciated with concrete crushing. Initiation of concrete crushing was for smaller increments of load. Because strain gauges were

86 ACI Structural Journal/January 2018


Fig. 10—Strains measured in selected Group L specimens.

Fig. 11—Strain distribution along splice in selected Group L specimens.


bonded at equal distances along the splice, the wider spacing T-heads at the ends of bars. Strain development was depen-
between curves at peak load observed for strains near the dent on splice length and presence of transverse reinforce-
end of the splice give evidence of the nonlinear development ment within the splice. Each graph in Fig. 12 presents six
of strain along the bar (refer also to Fig. 11(a)). curves, two for each of the three instrumented bars depicting
Measured strains in Specimen L-5-29 showed similar strains measured at two locations (end of splice and adjacent
trends as those discussed previously. Curves a, b, and c to T-head).
shown in Fig. 10(b), corresponding to the first three strain Strain values developed in specimens containing trans-
gauges starting from the end of the bar, exhibit a sudden verse reinforcement (T-5-8R and T-5-4R) within the splice
decrease at a load of approximately 17 kip (432 kN). The region were much higher than their companion specimens
decrease in strain near the end of the bar occurred at a load (T-5-8 and T-5-4). The maximum strain developed in spec-
corresponding to the formation of the first splitting crack imens without transverse reinforcement was between 0.002
along the splice that apparently caused the bars near the and 0.003 and, interestingly, sometimes occurred near the
splice end to slip. At higher loads, strains measured in these T-head. The change in strain from the T-head (triangles) to
three gauges no longer increased at the same rate as prior the end of the splice (circles) in these specimens was not
to formation of splitting cracks, further supporting the argu- substantial (Fig. 12(a) and (b)). The presence of transverse
ment for bar slip. reinforcement in Specimens T-5-8-R and T-5-4R resulted in
Strain distributions along the splice for four different load larger strains developing in each bar, particularly at the end
levels for the same two specimens discussed previously are of the splice region instead of near the T-head. Strain values
presented in Fig. 11. Strain variation along the splice for the in these specimens reached between 0.004 and 0.0045. The
two lowest load levels presented (cracking [Pcr] and 50% of increase in strain along each bar between the T-head and the
yield [Py]) was not as pronounced as for the two highest load splice end is also noted (Fig. 12(c) and (d)). The observa-
levels (Py and Pmax). Strains increased with distance from tions on strain behavior of specimens in Group T point to
the end of the spliced bar as would be expected. A higher differences in the force-transfer mechanism between adja-
strain gradient was observed in the region farthest from the cent lapped bars in specimens with and without transverse
bar end, but the strain values varied significantly in the two reinforcement along the splice.
cases presented. This result is not surprising because the To better understand the differences in force-transfer mech-
value of measured strain is strongly affected by strain gauge anisms between specimens in Group T, strut-and-tie models
location with respect to the location of cracks. were constructed for Specimens T-5-4 and T-5-4R (Fig. 13).
In specimens of Group T, the development of bar strains Solid lines represent ties and dashed lines indicate struts; the
along spliced bars was strongly influenced by the presence of force in each element of the models is expressed as a fraction

ACI Structural Journal/January 2018 87


Fig. 12—Comparison of measured bar strains in Group T specimens.
of the force in the exterior longitudinal bar Text. In Specimen
T-5-4, force transfer occurs primarily through the formation
of a diagonal strut between T-heads of two adjacent bars.
The tension force in longitudinal bars of this model is nearly
constant throughout the splice length consistent with the obser-
vation that measured strains next to the head and at the end of
the splice remained approximately constant in this specimen
(refer to Fig. 12(b)). The force in interior longitudinal bars of
the model is three times higher than the force in exterior bars.
This value is larger than the factor of 2.3 between interior and
exterior strains measured during the test, but the model serves
to illustrate the force-transfer mechanism that was observed.
The strut-and-tie model illustrating the force-transfer mecha-
nism of Specimen T-5-4R includes an intermediate tie corre-
sponding to the transverse reinforcement within the splice.
Presence of the middle tie increases the concrete strut inclina-
tion and allows the force in longitudinal ties to change within
the splice region, as was evidenced by changes in strains Fig. 13—Strut-and-tie models for Group T specimens
measured along bars during the experiments (compare strain containing 4 in. (102 mm) splice.
at head with strain at end of splice [Fig. 12(a)]). The force
ratio between interior longitudinal ties and exterior longitu- reinforcement. This assumption allowed determining the
dinal ties was 1.67 in this case, less than the ratio observed in fraction of the calculated flexural strength that each spec-
Specimen T-5-4. This force difference between interior and imen was able to reach depending on its splice length.
exterior bars was also noted in the strain measurements taken The moment-curvature response of each specimen was
during the test, as can be observed in Fig. 12(d). estimated using measured material properties in combina-
tion with the following assumptions. Concrete was consid-
Comparison of measured moments with estimated ered unconfined and modeled using the Mander et al. (1988)
flexural strength uniaxial stress-strain model with an unconfined strength
For comparison with experimentally determined moments, determined from the average measured compressive
flexural strength of all specimens was estimated using two strength of companion concrete cylinders. Tensile longitu-
procedures: a moment curvature analysis approach and a dinal reinforcement was modeled using Eq. (1a) for No. 4
simplified approach. Regardless of the procedure used, the (13 mm) and No. 5 (16 mm) bars, and Eq. (1b) for No. 6
longitudinal reinforcement was assumed to be continuous (19 mm) bars. The moment-curvature curve of each spec-
even in specimens containing lapped splices or T-headed imen was used to determine the flexural strength associated
with concrete crushing in compression after yielding of the
88 ACI Structural Journal/January 2018
tension reinforcement. Table 2 lists the calculated flexural lapped splices (Group L) were evaluated using the develop-
strength Mcalc and the ratio between the moment measured ment length models proposed by Hosny et al. (2012). Bar
during the test and the calculated flexural strength based on stresses of specimens in Group T were evaluated using ACI
moment-curvature analysis. The Mtest/Mcalc ratio for spec- 318-14 Section 25.4.4.2, which is based on the research by
imens with continuous reinforcement (Group C) ranges Thompson et al. (2006).
between 0.94 and 1.04. These ratios are less than 1.00 for Hosny et al. (2012) proposed a development length equa-
all specimens containing splices, as would be anticipated, tion for high-strength reinforcement conforming to ASTM
particularly for elements containing short splices. A1035/A1035M for unconfined conditions. The equation
An approximate approach was also followed to determine proposed was based on the results of 66 beams containing
flexural strength of the specimens as a way of computing a splices without transverse reinforcement along the splice.
reasonable estimate of strength without the need to conduct This equation was subsequently used to determine stresses
a moment-curvature analysis. Being a sectional model, this for comparison with test results from a much larger number
method also assumed that the reinforcement was continuous of tests (a total of 213 beams and slabs) to determine its
in all specimens, so flexural strength was reached when validity for other types of reinforcement. The resulting
concrete crushed in compression after bar yielding. Concrete development length equation of bars in tension without
stresses within the compression zone were represented using confining reinforcement was
the equivalent stress block described in ACI 318-14, Section
22.2.2.4. Tensile strength of concrete was neglected and ld (0.001 f s ) 2
perfect bond between concrete and tension reinforcement = (3a)
db 2 cmin f c′
was assumed. The stress-strain response of longitudinal
reinforcement was represented using a bilinear approxi-
mation with properties as shown in Fig. 3(b). To compute Solving for bar stress gives
flexural strength, the neutral axis depth was determined by
iteration until internal force equilibrium between tensile and 1000 f c′1/ 4 ld cmin
fs = (3b)
compressive forces was achieved within a reasonable toler- db
ance. Computed flexural strengths of all specimens using this
where cmin is the minimum concrete cover; fs is the bar stress;
simplified approach are listed in Table 2. Flexural strengths
db is the spliced bar diameter; and fc′ (in psi) represents the
calculated using the simplified method described in this
concrete compressive strength.
paper were not significantly different from those obtained
The tension development length for T-headed bars with
using the more rigorous moment-curvature approach. The
a nominal yield stress not exceeding 60,000 psi (420 MPa)
maximum difference was approximately 10% for L-4 spec-
and concrete compressive strength not exceeding 6000 psi
imens between the two approaches; differences in calcu-
(42 MPa), ACI 318-14, Section 25.4.4.2, requires that
lated strength of other specimen groups were below 5%. A
comparison between calculated and measured bar stresses
ld 0.016 f y
was then conducted using the simplified approach for spec- = (4a)
imens with continuous reinforcement (Group C). As shown db f c′
in Table 2, computed stresses compared favorably with those
determined during the tests by entering the bar stress-strain Replacing development length, ld, for splice length, ls, and
relation using average strains measured within the maximum the yield stress, fy, for bar stress fs, an expression for bar
moment region. stress can be used to compare with stress values measured
during the tests
EVALUATION OF LONGITUDINAL BAR STRESSES
Several existing equations to compute development length
ls f c′ 1
of bars in concrete were used to evaluate their applicability fs = ⋅ (4b)
to ASTM A1035/A1035M bars. All these equations were 0.016 db
developed from statistical analyses of test data that included, Bar stresses in Group L specimens were estimated using
in most cases, beam or slab tests with most of the reinforce- Eq. (3b) while those in Group T specimens were estimated
ment having a nominal yield stress of up to 60 ksi (420 MPa). using Eq. (4b), as listed in Table 2 under the column corre-
Only the development length equation proposed by Hosny sponding to fs,calc. Bar stresses in Group C and the Specimen
et al. (2012) included tests that had ASTM A1035/A1035M CT were found as part of the simplified flexural strength
bars. The applicability of the development length equations calculation described earlier in this paper, assuming a
evaluated by Hosny et al. is also limited to a maximum bilinear approximation for the stress-strain relation of the
concrete compressive strength of 10,000 psi (69 MPa), longitudinal reinforcement. Measured bar stresses in Groups
because higher concrete strength was not used in the beams L and C exceeded calculated values in most cases except for
reported by the researchers. three specimens (CT, C-5-M, and C-6-M). This result indi-
Many research studies have been devoted to develop- cates that the procedures used to estimate bar stresses that
ment length and lap splices of straight bars in tension. can then be used to estimate flexural strength were found
Fewer studies have concentrated on studying behavior of to be conservative. Measured (test) bar stresses in Group T
T-headed spliced reinforcement. Only specimens containing specimens were also higher than those calculated using

ACI Structural Journal/January 2018 89


Eq. (4b), except for Specimen T-5-8, where the measured forcement; Joint ACI-ASCE Committee 445, Shear and Torsion; and ACI
Subcommittee 318-C, Safety, Serviceability, and Analysis (Structural
stresses were 2% lower than estimated. In the majority of Concrete Building Code).
Group T specimens, the ratio of test-to-calculated stress was
much higher than 1.0, indicating that Eq. (4b) was very conser- Jeffrey Messier is a Project Manager with Schaefer in Phoenix, AZ. He
received his BS from the University of Massachusetts Dartmouth, Dart-
vative, in some instances by over 100%. The largest ftest/fcalc mouth, MA, in 2002 and his MS in civil engineering (structures) in 2004.
values correspond to specimens containing transverse rein-
forcement within the splice region of specimens in Group T. Sean W. Peterfreund is an Associate with McMillen Jacobs Associates,
Seattle, WA. He received his BS from the University of Massachusetts
This result differs from the findings of Thompson et al. (2006), Amherst in 2003 and his MS from the University of Michigan, Ann Arbor,
who showed that presence of transverse reinforcement had little MI, in 2004.
or no influence on splice strength of T-headed reinforcement.
ACI 318-14 provisions for T-headed reinforcement are based on ACKNOWLEDGMENTS
these researchers’ findings (Eq. (4a)). Larger-diameter T-heads The Commonwealth College at the University of Massachusetts provided
funds for the third author’s honors research. The authors would like to thank
were used in the tests presented in this paper, causing them to MMFX Technologies Corporation for donating the ASTM A1035/A1035M
be in contact with adjacent bars. This may have constrained reinforcement subject of this research, Erico for fabricating and providing
lateral spread of compression stresses generated in front of mechanical couplers and terminators, and Barker Steel for providing trans-
verse and compression reinforcement used in the specimens.
the head, promoting efficiency of transverse reinforcement in
increasing the strength of the splice.
NOTATION
c = spacing or cover dimension = cmin + db/2, in.
CONCLUSIONS cmax = maximum concrete cover (side or bottom), in.
Splice strength tests of straight and T-headed high-strength cmin = minimum concrete cover (side or bottom), in.
db = bar diameter, in.
reinforcement (ASTM A1035/A1035M) were conducted to fc′ = concrete compressive strength, psi
provide information on behavior, to understand the force- fs = reinforcing bar stress, psi
transfer mechanisms within the splice region, and to deter- fy = yield stress of reinforcing bar, psi
Ktr = transverse reinforcement index = (0.52trtdAtr/sn)fc′
mine the influence of splice length in stress developed in ld = development length of reinforcing bar, in.
longitudinal bars. The following conclusions may be drawn ls = splice length of reinforcing bar, in.
as a result of these series of tests: Rr = relative rib area of the reinforcement ≈ 0.073
td = 0.78db + 0.22
1. The specimens tested in this research containing straight tr = 9.6Rr + 0.28 ≤ 1.72
bar spliced high-strength reinforcement failed by splitting of εs = strain in reinforcing bar
concrete. This failure mode could potentially be eliminated ϕ = strength reduction factor
ω = (0.1 cmax/cmin + 0.9) ≤ 1.25
if transverse reinforcement were present in the splice region.
2. Straight bar lap splices involving high-strength reinforce-
ment may become impractically long if the design goal is to REFERENCES
ACI Committee 318, 2014, “Building Code Requirements for Struc-
promote a ductile failure involving reinforcing bar yielding tural Concrete (ACI 318-14) and Commentary (ACI 318R-14),” American
followed by concrete crushing in compression, thereby Concrete Institute, Farmington Hills, MI, 519 pp.
ACI Innovation Task Group 6, 2010, “Design Guide for the Use of ASTM
avoiding splitting failures. Again, the presence of transverse A1035/A1035M Grade 100 (690) Steel Bars for Structural Concrete (ACI
reinforcement within the splice length would help promote ITG-6R-10),” American Concrete Institute, Farmington Hills, MI, 90 pp.
ductility by eliminating or delaying splitting failures. Applied Technology Council, 2014, “Roadmap for the Use of High-
Strength Reinforcement in Reinforced Concrete Design (ATC-115),”
3. Flexural strength of members containing high-strength Redwood City, CA, 194 pp.
reinforcement without a well-defined yield plateau can be Hosny, A.; Seliem, H. M.; Rizkalla, S.; and Zia, P., 2012, “Development
estimated with sufficient accuracy by adopting simplified Length of Unconfined Conventional and High-Strength Steel Reinforcing
Bars,” ACI Structural Journal, V. 109, No. 5, Sept.-Oct., pp. 655-664.
models for concrete in compression (stress block) and repre- Joint ACI-ASCE Committee 408, 2003, “Bond and Development of
senting the stress-strain relation of steel by a bilinear relation. Straight Reinforcing Bars in Tension (ACI 408R-03),” American Concrete
4. Couplers may be used to connect high-strength rein- Institute, Farmington Hills, MI, 49 pp.
Mander, J. B.; Priestley, M. J. N.; and Park, R., 1988, “Theoret-
forcing bars and promote ductile flexural failure associ- ical Stress-Strain Model for Confined Concrete,” Journal of Struc-
ated with bar yielding in tension and concrete crushing in tural Engineering, ASCE, V. 114, No. 8, pp. 1804-1826. doi: 10.1061/
compression of elements containing this reinforcement. (ASCE)0733-9445(1988)114:8(1804)
Mast, R. F.; Dawood, M.; Rizkalla, S. H.; and Zia, P., 2008, “Flexural
5. The presence of transverse reinforcement within the Strength Design of Concrete Beams Reinforced with High-Strength Steel
lap splice region of T-headed bars changes the force-transfer Bars,” ACI Structural Journal, V. 105, No. 4, July-Aug., pp. 570-577.
mechanism between adjacent bars and leads to the develop- Seliem, H. A.; Hosny, A.; Rizkalla, S.; Zia, P.; Briggs, M.; Miller, S.;
Darwin, D.; Browning, J.; Glass, G. M.; Hoyt, K.; Donnelly, K.; and Jirsa,
ment of higher stresses along the splice and higher member J. O., 2009, “Bond Characteristics of ASTM A1035 Steel Reinforcing
deformation capacity for equal lap lengths. Bars,” ACI Structural Journal, V. 106, No. 4, July-Aug., pp. 530-539.
Shahrooz, B. M.; Reis, J. M.; Wells, E. L.; Miller, R. A.; Harries, K. A.;
and Russell, H. G., 2014, “Flexural Members with High-Strength Reinforce-
AUTHOR BIOS ment: Behavior and Code Implications,” Journal of Bridge Engineering,
Sergio F. Breña, FACI, is a Professor at the University of Massachusetts
ASCE, V. 19, No. 5, pp. 1-7. doi: 10.1061/(ASCE)BE.1943-5592.0000571
Amherst, Amherst, MA. He received his MS and PhD from the University
Thompson, M. K.; Ledesma, A.; Jirsa, J. O.; and Breen, J. E., 2006, “Lap
of Texas at Austin, Austin, TX. He is a member of ACI Committees 369,
Splices Anchored by Headed Bars,” ACI Structural Journal, V. 103, No. 2,
Seismic Repair and Rehabilitation; 374, Performance-Based Seismic
Mar.-Apr., pp. 271-279.
Design of Concrete Buildings; 440, Fiber-Reinforced Polymer Rein-

90 ACI Structural Journal/January 2018


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 115-S07

End-Region Behavior of Pretensioned I-Girders Employing


0.7 in. (17.8 mm) Strands
by J. Salazar, H. Yousefpour, R. Alirezaei Abyaneh, H. Kim, A. Katz, T. Hrynyk, and O. Bayrak
This paper presents an experimental study on transfer length, tions.4 In ACI 318-14 provisions,5 ltr is taken as (fse/3000)db,
end-region cracking, and transverse end-region stresses in preten- in which fse is the effective stress in the strands after prestress
sioned concrete I-girders fabricated using 0.7 in. (17.8 mm) diam- losses in psi, or (fse/21)db if fse is in MPa. Therefore, on the
eter strands on a 2 x 2 in. (51 x 51 mm) grid. The full-scale specimens basis of these provisions, an increase in the strand diameter
consisted of two Tx46 and two Tx70 girders that were fabricated in
is expected to result in an increase in the transfer length.
a controlled laboratory environment using different strand patterns
However, studies supporting the aformentioned provisions
and concrete release strengths. The detailing for mild-steel rein-
forcement was done according to the current practice in Texas were done considering strands with diameters that were less
for girders with smaller-diameter strands. The measured 24-hour than or equal to 0.6 in. (15.2 mm),6 and the applicability of
transfer lengths from the specimens exceeded estimates by both such provisions to larger-diameter 0.7 in. (17.8 mm) strands
AASHTO LRFD and ACI 318-14 provisions. The observed crack has yet to be verified.
widths in the specimens within 28 days after prestress transfer were To minimize the impact of transitioning to larger-diameter
generally limited to 0.007 in. (0.18 mm), indicating satisfactory 0.7 in. (17.8 mm) strands on fabrication facilities, it is of
performance for exposure to deicing chemicals according to ACI considerable interest to use these strands on the 2 x 2 in. (51 x
224R guidelines. However, noticeably greater transverse forces 51 mm) grid that is typically used for 0.5 and 0.6 in. (12.7 and
were observed in the end regions of the specimens compared to the 15.2 mm) strands. However, incorporating 0.7 in. (17.8 mm)
resistance required by AASHTO LRFD specifications.
strands in this configuration is believed to increase the stresses
Keywords: 0.7 in. strands; bursting stresses; end-region cracking; preten- that develop within the end region of pretensioned girders
sioned; spalling stresses; splitting resistance; transfer length. and, as such, may lead to end-region damage that is not found
in girders constructed with smaller-diameter strands.
INTRODUCTION Transfer of prestress within girder end regions results in
Pretensioned concrete bridge girders are currently fabri- transverse stresses that are generally categorized as bursting
cated using 0.5 or 0.6 in. (12.7 or 15.2 mm) diameter and spalling stresses.7 Bursting stresses are mostly due to
prestressing strands. In recent years, however, it has become Poisson’s effect at the time of release. When the strands
of interest to use 0.7 in. (17.8 mm) diameter seven-wire are released, they tend to return to their original diameter,
strands in pretensioned elements. These larger-diameter causing radial stresses in the surrounding concrete.8 These
strands have so far been used primarily for mining applica- stresses are typically located in the bottom flange of preten-
tions and cable bridges.1 In addition to reducing the number sioned girders and can result in longitudinal crack develop-
of required strands, the increase in the strand diameter will ment along the strands. Spalling stresses are due to compat-
allow for a greater concentration of steel to exist closer to the ibility of the strains near the end face of the girders and are
tensile face of the cross section. Therefore, nominal flexural dependent on the eccentricity of the strands. These stresses
and shear capacities are expected to be greater in members might result in horizontal or inclined cracks that are close to
with larger-diameter strands. Such capacity increases may the end face of the member but are typically located at some
offer improvements in attainable span capabilities, slen- distance from the elevation of the strands.
derness of bridge superstructure, and transverse spacing To alleviate concerns regarding the serviceability and
between girders.2 durability of girders, appropriate reinforcement detailing is
In pretensioned concrete elements, the stress in the strand essential for controlling the width of potential cracks due
increases from zero at the end face of the member to the full to bursting and spalling stresses. Confining reinforcement
prestress level over a distance referred to as “transfer length.” surrounding the strands and transverse reinforcement in the
A longer transfer length results in a more gradual develop- web, near the ends of the member, should be used to resist
ment of prestress with distance from the end face and, there- bursting and spalling stresses, respectively. The amount of
fore, reduces the end-region damage at the time of prestress such reinforcement in current detailing practice has been
transfer. However, reduced prestress levels developed within primarily based on experience or from the results of previous
the transfer length might negatively affect the strength of experimental studies on members with 0.5 and 0.6 in. (12.7
pretensioned girders, especially under shear-critical loading and 15.2 mm) diameter strands. Therefore, the suitability of
conditions.3 Therefore, the transfer length ltr is a critical ACI Structural Journal, V. 115, No. 1, January 2018.
design parameter. In most design provisions, ltr is considered MS No. S-2016-309.R2, doi: 10.14359/51700783, received February 13, 2017, and
reviewed under Institute publication policies. Copyright © 2018, American Concrete
primarily a function of strand diameter db. A value of 60db Institute. All rights reserved, including the making of copies unless permission is
is used in AASHTO LRFD 2014 Bridge Design Specifica- obtained from the copyright proprietors. Pertinent discussion including author’s
closure, if any, will be published ten months from this journal’s date if the discussion
is received within four months of the paper’s print publication.

ACI Structural Journal/January 2018 91


Table 1—Summary of previous studies on specimens with 0.7 in. (17.8 mm) strands
Reported transfer
lengths, in.
Release
Source Specimen type Nspec Strand layout fpi/fpu fci′, ksi method ltr,initial ltr,t t, days Notes
Tadros and
96 x 7 x 7 in. 4 One strand 0.75 6.0 G 26 to 28 26 to 28 28 Small-scale prism
Morcous1
One or two
Dang et al.9 216 x 6.5 x 12 in. 16 0.75 5.9 to 9.2 G 23 to 28 26 to 31 28 Small-scale prism
strands
Vadivelu10 AASHTO Type I 1 2.0 x 2.0 in. 0.75 10.0 F 21 to 22 — — —
T-girders 6 2.0 x 2.0 in. 0.75 9.0 G 19 to 25 21 to 26 28 —
Tadros and
NU900 1 2.0 x 2.0 in. 0.66 12.0 F 26 — — UHPC
Morcous1
BDT 1 2.0 x 2.0 in. 0.60 12.0 F 14 to 20 17 to 21 14 UHPC
Morcous et al. 11
NU900 1 2.2 x 2.3 in. 0.75 6.7 F 35 — — —
Morcous et al. 12
NU1350 2 *
2.0 x 2.0 in. 0.75 6.0 to 10.0 F 32 36 14 Bridge application
*
Morcous et al.12 reported transfer length measurements from two girders out of 20 fabricated girders.
Notes: Nspec is number of specimens tested; fpi is jacking stress; fpu is ultimate strength of prestressing steel; fci′ is compressive release strength; ltr,initial is transfer length immediately
after prestress transfer; ltr,t is transfer length at t; G is gradual release; F is flame cut; and BDT is bridge double tee; 1 in. = 25.4 mm; 1 ksi = 6.9 MPa.

such detailing for girders employing 0.7 in. (17.8 mm) diam- RESEARCH SIGNIFICANCE
eter strands needs to be investigated. The use of 0.7 in. (17.8 mm) diameter prestressing strands
Relatively few experimental studies have addressed might influence the serviceability and strength of preten-
prestress transfer or end-region behavior in girders sioned girders due to increased transfer length and greater
employing 0.7 in. (17.8 mm) diameter strands. A summary end-region stresses. The experimental study presented
of those studies is provided in Table 1. As can be seen in in this paper is a major contribution to the knowledge
this table, some of the studies were conducted on small- regarding the performance of full-scale girders employing
scale specimens that were reinforced with only one strand. these larger-diameter strands at the time of prestress transfer,
Such studies do not provide a realistic representation of the in which different strand patterns, girder cross section sizes,
boundary conditions or of the interaction between adjacent and concrete release strengths have been investigated. More-
strands, and the applicability of results from such studies to over, this study presents the first set of data on bursting and
full-scale girders is questionable. spalling stresses and progression of end-region cracking
Full-scale specimens have been used in a few studies to within the first few weeks after the fabrication of girders
evaluate transfer length and constructability issues for girders with 0.7 in. (17.8 mm) diameter strands.
employing 0.7 in. (17.8 mm) diameter strands. The transfer
length, which was determined using mechanical measure- EXPERIMENTAL PROGRAM
ments of surface strains in all these studies, was reported to Four pretensioned Texas Bulb-Tee girders, commonly
be noticeably less than the estimates from ACI 318-14 and referred to as Tx-girders, were fabricated using straight
AASHTO LRFD 2014. Very little information was reported 0.7 in. (17.8 mm) diameter seven-wire strands that were
in any of these studies regarding the cracks developed within located on a 2 x 2 in. (51 x 51 mm) grid. These specimens
the end regions of the specimens. Moreover, the magnitude of included two Tx46 and two Tx70 girders, each with a length
bursting and spalling stresses was not measured in any of the of 30 ft (9 m) and the cross-sectional geometry shown in
studies, providing little insight into any need to modify the Fig. 1. The cross-sectional properties of the specimens are
end-region detailing for use in girders with 0.7 in. (17.8 mm) listed in Table 2.
diameter strands. Furthermore, in most of these studies, the A summary of the design parameters for the specimens
prestress transfer was performed at a concrete release strength is shown in Table 3. While the Tx46 specimens represent
considerably greater than what is commonly used in practice. the mid-sized bulb-tee cross sections that are widely used
Therefore, the behavior of the specimens and the observed in Texas, the Tx70 specimens represent the deepest girders
transfer lengths may not be indicative of the performance of among this family of precast sections and have been shown
actual pretensioned girders used in the field. to likely experience the greatest benefit from using larger-
This paper presents an experimental research program on diameter strands.2 Different strand patterns were used among
full-scale pretensioned I-girders that were fabricated using the specimens, as shown in Fig. 2. The strands in Tx46-I
0.7 in. (17.8 mm) diameter strands in the controlled laboratory and the Tx70 specimens were placed conventionally—that
environment. The compressive strength of concrete at the time is, at the greatest possible eccentricity within the specimen
of release ranged from 5.2 to 8.3 ksi (35.8 to 57.2 MPa), and cross section—to generate the maximum spalling stresses
the girders were extensively instrumented to monitor bursting within the specimen end regions. In Tx46-II, the strands
stresses, spalling stresses, and transfer length. were concentrated near the centroid of the cross section to
represent critical conditions for bursting stresses.

92 ACI Structural Journal/January 2018


Fig. 1—Standard detailing for mild-steel reinforcement in Tx-girders.13 (Note: Clear distance between first pair of S-bars and
end face: 1.5 in. [38 mm]; 1 in. = 25.4 mm.)

Table 2—Cross-sectional properties of Tx46 and Table 3—Specimen design parameters


Tx70 girders13 Tx46-I Tx46-II Tx70-I Tx70-II
Property Tx46 Tx70 Max. Max. Max.
Gross cross-sectional area, in.2 761 966 Design objective eccentricity Max. Pi eccentricity eccentricity

Distance from bottom fiber to center of gravity of Design fci′, ksi 5.5 5.2 5.5 7.8
20.1 31.9
girder, in. Top 4 4 4 4
No. of
Moment of inertia around x-axis, in.4 198,089 628,747 strands Bottom 24 30 28 42
Moment of inertia around y-axis, in.4 46,478 57,579 Top strands 44.0 44.0 68.0 68.0
Weight, lb/ft 819 1040 yp, in. Bottom
3.3 10.4 3.5 4.5
Notes: 1 in. = 25.4 mm; 1 in.2 = 645 mm2; 1 in.4 = 416,000 mm4; 1 lb/ft = 1.5 kg/m.
strands
Top strands 157.5 202.5 110.0 202.5
The specimens were designed based on AASHTO LRFD
fpi, ksi
Specifications.4 According to these specifications, the allow- Bottom
202.5 202.5 202.5 202.5
able stresses at the time of prestress transfer are 0.65fci′ in strands
compression and k√fci′ in tension, where fci′ is the compres- σpredicted, Top fiber 0.23 (T) 1.13 (C) 0.55 (T) 0.55 (T)
sive strength of concrete at the time of prestress transfer, ksi Bottom fiber 3.57 (C) 3.40 (C) 3.53 (C) 5.04 (C)
and k is 0.24 or 0.63, if fci′ is in ksi or MPa, respectively. In
Notes: Pi is initial prestressing force; yp is distance from bottom fiber of girder to
addition to the strands in the bottom flange, all specimens centroid of strands; fpi is jacking stress; σpredicted is predicted concrete stress after
included four 0.7 in. (17.8 mm) diameter strands within the prestress transfer; (C) is compression; (T) is tension;. 1 ksi = 6.9 MPa; 1 in. =
top flange to control the stresses within the cross section at 25.4 mm.

ACI Structural Journal/January 2018 93


Fig. 2—Strand patterns within bottom flange of specimens.
diameter, Grade 270 strands were procured from a commer-
cial prestressing steel manufacturer. Reusable strand chucks
were used for anchoring the 0.7 in. (17.8 mm) diameter
strands at each end, which had an external diameter of 2 in.
(51 mm). To remove the slack and ensure uniform stressing
of the strands, each strand was individually stressed to 2 kip
(9 kN), after which the main hydraulic rams were used to
“gang-stress” the strands—that is, stress all strands simul-
taneously. Gang-stressing of the strands was carried out
in a minimum of 10 increments. The prestress level was
controlled through measurements of hydraulic pressure in
the rams using pressure transducers. The elongation of the
strands was measured using a series of linear potentiome-
ters at each end of the prestressing facility and verified to be
within 5% of the calculated value after each increment. Once
Fig. 3—Prestressing facility at FSEL. the strands were stressed to the desired level, installation of
transverse reinforcement was carried out.
the time of prestress transfer. The stress level in these strands
Typically, each specimen contained 90 electrical resistance
was variable among the specimens to optimize the capacity
strain gauges (SGs) on the strands to measure the transfer
of the specimens and maximize the number of strands that
length. At each end of the specimen, four strands were
could be accommodated in the specimens. The end-region
instrumented, as shown in Fig. 4. Among the instrumented
and shear reinforcement detailing for the specimens followed
strands, the SGs were installed at two intervals—12.0 in.
the standard drawings developed by the Texas Department
(305 mm) and 6.0 in. (152 mm)—to examine any potential
of Transportation (TxDOT)13 for girders with 0.5 and 0.6 in.
changes in transfer length due to the presence of SGs. The
(12.7 and 15.2 mm) diameter strands, as shown in Fig. 1.
SGs were continued up to a distance of 60 in. (1.52 m) from
Specimens Tx46-I, Tx46-II, and Tx70-I were designed
the end face of the specimens. One SG was also installed
assuming concrete release strengths that are typically used
outside the specimen on each of the instrumented strands
for prestressed concrete superstructures in Texas. However,
to serve as a reference measurement representing the stress-
a greater concrete release strength was considered for the
free strain condition of the strand after prestress transfer. All
design of Tx70-II to investigate the behavior of girders fabri-
strand SGs were installed on the helical wires. An example
cated with a greater number of 0.7 in. (17.8 mm) diameter
of strain gauge application on a strand is shown in Fig. 5(a).
strands. As a result, Tx70-II could accommodate an addi-
Each specimen also contained a considerable number of
tional row of strands compared to Tx70-I.
SGs on the transverse reinforcement to determine bursting
The specimens were fabricated using the prestressing
and spalling stresses as well as strains at the time of future
facility at Ferguson Structural Engineering Labora-
structural tests. At a minimum, the first 15 stirrups (R-bars
tory (FSEL) (Fig. 3), which is designed to accommodate
in Fig. 1) from each end of the specimen were instrumented.
prestressing forces up to 2500 kip (11,120 kN). Steel bulk-
Three SGs were installed on each instrumented stirrup, as
heads measuring 12 in. (305 mm) thick are used at each
shown in Fig. 4. The stirrup leg that contained two SGs was
end of the prestressing facility to anchor the strands. The
alternated from one stirrup to the next. The initial installa-
bulkhead at one end is fixed, while the bulkhead at the other
tion of a SG on one of the stirrups, prior to applying protec-
end is supported by four 800 kip (3500 kN) hydraulic rams,
tive layers, is shown in Fig. 5(b).
which are extended for stressing the strands and retracted at
The specimens were also instrumented using three
the time of release. These two ends are herein referred to as
vibrating wire strain gauges (VWGs) that were embedded at
the dead and live ends, respectively. Two adjustable cross
the midspan of the specimen (Fig. 5(c)). Measurements from
beams are also used within the prestressing facility, making
VWGs were used to develop a strain profile at the midspan
it possible to apply prestressing forces of up to 300 kip
section and to estimate prestress losses due to elastic short-
(1330 kN) to top strands in different types of cross sections.
ening, shrinkage, and creep over the life of the specimen.
Fabrication of each specimen began by placing the 0.7 in.
To measure the hydration temperatures in the concrete, six
(17.8 mm) diameter strands in the prestressing facility and
thermocouples were placed in a section located within 2 ft
installing the anchorage devices. The 0.7 in. (17.8 mm)
(0.60 m) from the end face of the specimens. As shown in

94 ACI Structural Journal/January 2018


Fig. 4—SG locations for monitoring transfer length and transverse reinforcement stresses. (Note: 1 in. = 25.4 mm.)
Table 4—Summary of measured mechanical
properties
Test
Property method Tx46-I Tx46-II Tx70-I Tx70-II
ASTM
fci′, ksi 5.7 5.2 6.5 8.3
C3914
ASTM
Concrete Eci, ksi — 4940 4490 4900
C46915
ASTM
f ′c,28, ksi 7.3 7.0 10.7 12.7
C3914
fy, ksi 60.7 72.2
No. 4 bars
fu, ksi ASTM 99.3 111.6
fy, ksi A37016 74.0 67.7
No. 6 bars
fu, ksi 114.7 106.7
Ep, ksi 27,810
ASTM
Strands fpy, ksi 232.0
A106117
fpu, ksi 276.1

Notes: fci′ and Eci are compressive strength and modulus of elasticity of concrete at
prestress transfer, respectively; fc,28′ is 28-day compressive strength of concrete;
1 ksi = 6.9 MPa.

precast manufacturing plant and transported to FSEL. In all


Fig. 5—Instrumentation application. cases, a combination of internal and external vibration was
Fig. 5(d), these thermocouples were distributed within the used to ensure satisfactory consolidation of concrete.
cross section to capture the variability of concrete curing Forty-eight 4 x 8 in. (100 x 200 mm) match-curing concrete
temperatures within the end region of the girders. cylinders were cast using the concrete comprising each spec-
The properties of concrete mixtures used within the test imen. These cylinders were connected to a relay system that
program are provided in Appendix A.* Type III portland maintained the same temperatures within the cylinders as
cement and crushed limestone with a maximum aggregate those measured from the embedded thermocouples. There-
size of 3/8 in. (10 mm) were used in all mixtures. The clear fore, eight cylinders were match-cured based on each of
spacing between the prestressing strands was approximately the six thermocouples shown in Fig. 5(d), which made it
1.3 in. (33.0 mm), which exceeds 1.33 times the maximum possible to capture the variability of concrete strength within
size of the aggregates, 1 in. (25.4 mm), and the strand the specimen cross section.
diameter. Therefore, the minimum spacing requirements of In the hours following the concrete placement, match-
ACI 318-14 and AASHTO LRFD 2014 were satisfied. cured specimens were periodically tested to identify the
The mixtures used for the Tx46 specimens were batched appropriate timing for formwork removal and subsequent
and mixed at FSEL. The concrete comprising the Tx70 prestress transfer. Prestress transfer commenced as soon as
specimens was batched and mixed by a nearby commercial the compressive strength of match-cured cylinders based
on all thermocouple measurements exceeded the desired
release strength. The prestressing strands were released
*
The Appendixes are available at www.concrete.org/publications in PDF format, through gradual retraction of hydraulic rams in 20 steps,
appended to the online version of the published paper. It is also available in hard copy which typically took 1 hour. As the strands were being
from ACI headquarters for a fee equal to the cost of reproduction plus handling at the
time of the request.
released, the compressive strength and modulus of elas-

ACI Structural Journal/January 2018 95


ticity of match-cured concrete specimens were measured to Table 5—Transfer length estimates from live and
obtain the mechanical properties at prestress transfer. The dead ends of specimens
average compressive strength and modulus of elasticity of Live end Dead end
the match-cured specimens at the time of prestress transfer
lplateau, ltr, lplateau, ltr,
are presented in Table 4.
in. in. n in. in. n
Along with the match-cured cylinders, additional cylin-
ders were prepared to measure mechanical properties of At Max 42 40 48 47
1 4
concrete at 28 days. Moreover, the mechanical properties of release Min. 42 40 42 40

Tx 46-I
the prestressing strands and the mild-steel reinforcing bars Max. 48 45 48 46
used in the construction of the girders were determined by 24 hour 1 3
Min. 48 45 42 41
performing ASTM-compliant tests at FSEL, which are also
presented in Table 4. Each mechanical property reported At Max. 42 35 48 41
4 4
in this table is an average value from a minimum of three release

Tx 46-II
Min. 36 29 36 29
samples tested by the research team. Max. 54 52 54 48
Data acquisition from the instrumentation commenced 24 hour 3 3
Min. 48 44 48 45
immediately prior to prestress transfer and was continued
for 24 hours after the end of the release operation. After the At Max. 42 39 42 41
4 3
release operation was completed, the specimens were care- release Min. 36 29 36 34

Tx 70-I
fully examined for end-region cracking. Max. 48 46 54 51
24 hour 3 2
Min. 36 31 36 35
RESULTS AND DISCUSSION
Transfer length At Max. 48 43 42 40
4 3
The transfer length was determined by comparing the data release
Tx 70-II Min. 36 34 36 34
obtained from the SGs before the release operation with those Max. 48 45 48 45
obtained immediately after release and 24 hours after release. 24 hour 2 3
Min. 48 45 48 41
To determine the transfer lengths, a modified version of the
95% average maximum strain (AMS) method introduced by At Max. 48 43 48 47
13 14
Summary

Russell and Burns6 was used. With the increase in distance release Min. 36 29 36 29
from the end face, the strains gradually increased from zero Max. 54 52 54 51
until reaching a plateau. For each end of each specimen, the 24 hour 9 11
Min. 36 31 36 35
strains in the plateaued region were averaged to determine
the average maximum strain. The transfer length is defined Notes: n is number of strands used for determining the transfer length; 1 in. = 25.4 mm.
as the distance at which the strain-versus-distance plot inter-
sects the 95% of AMS. The strains used in this procedure
were obtained from SGs that were installed on helical wires,
which are linearly correlated to the average axial strain in
the strand.18 Therefore, using these strains as opposed to the
average axial strain in the strand is not expected to affect the
transfer lengths.
Detailed information related to the evaluation of the transfer
lengths for the specimens is provided in Appendix B. Table 5
provides a summary of the distances corresponding to the start
of the plateau region, lplateau, and to the 95% AMS at dead and
live ends of each specimen. Because some of the strain gauges
did not function properly, determining the transfer length was
not possible for all instrumented strands. For each end region,
the number of strands from which a reliable transfer length Fig. 6—Comparison between measured transfer lengths and
could be determined is shown in the table. code predictions. (Note: 1 in. = 25.4 mm; 1 ksi = 6.9 MPa.)
In general, the transfer lengths obtained from live and which is consistent with the known effect of time-dependent
dead ends of each specimen were similar. This observation deformations of concrete on transfer length.19 At 24 hours
comes as no surprise because the gradual release of strands after release, the shortest and longest transfer lengths were
by hydraulic rams is expected to result in little difference recorded as 31 in. (790 mm) and 52 in. (1320 mm), obtained
between the live and dead ends of the specimens. Imme- from the live end of Tx70-I and the live end of Tx46-II,
diately after release, the shortest transfer length was 29 in. respectively.
(740 mm), which was obtained from both ends of Tx46-II. Figure 6 provides a comparison between the transfer
The longest transfer length at this time was 47 in. (1190 mm), lengths determined in this study with estimates from
which was found at the dead end of Tx46-I. AASHTO LRFD Specifications and ACI 318-14. The
Table 5 also shows a noticeable increase in the transfer vertical bars in this figure present the average transfer length
lengths measured within the first 24 hours after release, determined for each specimen at release and at 24 hours. The

96 ACI Structural Journal/January 2018


observed range between maximum and minimum transfer Table 6—Short- and long-term prestress losses
lengths is also illustrated using the vertical black lines. The Tx46-I Tx46-II Tx70-I Tx70-II
transfer length calculation in ACI 318-14 is dependent on
∆fES, ksi 21.5 18.9 24.1 30.7
effective stress in the strands after prestress losses and there-
fore varies among the specimens. The prestress loss due to t at final
131 40 39 28
elastic shortening, ∆fES, was estimated using data obtained measurement
from the VWGs and is shown in Table 6. ∆ftotal,t, ksi 45.1 35.6 38.7 45.9
Figure 6 shows that the transfer lengths immediately Note: 1 ksi = 6.9 MPa.
after release were generally shorter than those predicted by
ACI 318-14 and AASHTO LRFD provisions. However, after width observed. Similar figures reflecting the cracking
24 hours, the transfer length in all specimens, except Tx70-I, patterns on the other side of each specimen are included in
exceeded the estimates by both provisions. In Tx70-I, the Appendix C. In Fig. 7, the small circles are used to present
average 24-hour transfer length was very close to estimates the measured width of the cracks at each location at the
by both codes, but transfer lengths greater than code esti- time of final crack measurements. In the regions where no
mates were also observed. A greater increase in the transfer circles are shown, the crack width was less than or equal
lengths is anticipated for all specimens over time. However, to 0.004 in. (0.10 mm). The widths of the cracks in the first
monitoring the growth in transfer length after 24 hours was three specimens were measured using a crack comparator
not considered in this study. with a resolution of 0.002 in. (0.05 mm). For Tx70-II, a
The transfer lengths observed in this study were greater 7× magnifying loupe was used with reticles that provided
than those reported in previous studies noted in Table 1. In a crack measurement resolution of 0.0004 in. (0.01 mm).
those studies, the overestimation of transfer length by the However, for brevity and simplicity, crack widths shown for
design codes were deemed to be primarily due to the high Tx70-II are also categorized in a manner similar to that used
release strength of concrete, which is not explicitly consid- for the other specimens.
ered in code equations. However, observations from Fig. 6 The three specimens constructed with a conventional
do not support that reasoning. As can be seen in the figure, strand pattern revealed spalling cracks in their end regions.
the transfer lengths immediately after and 24 hours after As the prestressing force increased near the bottom fiber of
release were not correlated with the release strength, indi- these specimens, the spalling cracks extended further into
cating that the improved bond due to greater compressive the beam. Tx46-II, which was designed to accommodate
strength is not the only factor affecting the transfer length. the greatest prestressing force, showed bursting cracks that
The use of strain gauges in this study is believed to provide were primarily limited to the bottom flange. All specimens
a more precise picture of stress changes in the strands showed near-continuous cracking at the interface between
compared to the mechanical measurements of surface the web and the bottom flange within the end region. In all
strains. However, the presence of strain gauges could also specimens except Tx46-II, such cracking was also observed
potentially have a negative effect on the prestress transfer through the width of the web at the end faces. In Tx70-II,
within the instrumented strands. Thus, the transfer lengths cracks parallel to the outermost strands, which are indicative
from strands instrumented at 6 and 12 in. (150 and 300 mm) of bond-related damage, were also observed.
intervals were compared. The results showed that the The maximum crack width recorded over the course of
transfer lengths from strands with 6 in. (150 mm) spacing the entire test program was 0.008 in. (0.20 mm). This crack
were consistently greater than those with gauges placed width was observed immediately after prestress transfer in
at 12 in. (300 mm) spacing. However, the difference was the spalling cracks within the web of Tx46-I and Tx70-I.
generally limited to 6 in. (150 mm), which is the resolution Note that the crack widths in both specimens were recorded
of the estimated transfer lengths. Therefore, the presence of with a resolution of 0.002 in. (0.05 mm). The maximum crack
strain gauges does not seem to be the major contributor to widths in Tx46-II and Tx70-II were recorded as 0.004 in.
the greater transfer lengths measured in this study. (0.10 mm) and 0.007 in. (0.18 mm), respectively. In Tx46-I
and Tx70-I, the reported maximum width was observed only
Cracking patterns in an isolated length of few cracks. However, in Tx70-II,
All specimens were carefully inspected for cracks imme- crack widths between 0.006 and 0.007 in. (0.16 and
diately after prestress transfer. Moreover, a final survey of 0.18 mm) were measured at several locations. The greatest
the specimens for cracks was conducted at least 28 days after crack width was observed in the dead end of all specimens
the specimen was cast. For Tx70-II, in addition to the initial except Tx70-I. However, the overall difference in patterns or
and final measurements, cracks were also measured and widths of cracks between the two ends was not significant.
documented at ages of 7 and 13 days. The age of the speci- All specimens demonstrated noticeable changes in their
mens at the time of final crack measurements, t, is presented cracking conditions over time. These changes included
in Table 6. This table also includes the prestress losses in the growth in the length and width of cracks that were
bottom strands at the time of final measurements, ∆ftotal,t, as detected immediately after prestress transfer, as well as the
determined from VWGs. development of new cracks, especially in the Tx70 speci-
The measured cracking patterns of the specimens are mens. However, in most end regions, the widest crack did
presented in Fig. 7. This figure shows both end faces but not demonstrate noticeable growth between the time of
only the side of each specimen containing the greatest crack release and the time of final measurement. Periodic observa-

ACI Structural Journal/January 2018 97


Fig. 7—Crack patterns and widths in specimens. (Note: 1 ft = 0.30 m; 1 mm = 0.039 in.)
tions from Tx70-II also showed that most of the new cracks to ACI 224R, all girders comprising this test program met
formed within the first week after the prestress transfer. the conditions for use in exposure to deicing chemicals.
There is no globally accepted limit for the permissible 2. NCHRP Report 654, a comprehensive study on the
crack width within the end regions of pretensioned concrete acceptance criteria for the width of end-region cracks by
elements. To evaluate the performance of the specimens in Tadros et al.21 recommends that no action be taken for any
this test program, three references were used, as follows: end-region cracks that are 0.012 in. (0.30 mm) in width or
1. ACI 224R-01,20 a report by ACI Committee 224 on less. The observed crack widths in this study did not exceed
concrete cracking, provides general guidelines on acceptable the recommended limit. Therefore, no repair is required
crack widths in reinforced concrete flexural elements under according to these guidelines.
service loads. 3. TxDOT specifications for construction and maintenance
According to these guidelines, the “reasonable” crack of highways, streets, and bridges22 require corrective action
width is 0.007 in. (0.18 mm) for elements exposed to deicing if cracks exceeding 0.005 in. (0.13 mm) form within the end
chemicals and 0.012 in. (0.30 mm) for elements exposed to regions of I-girders. While the specimens did not satisfy this
humidity, moist air, and soil. Table 4.1 in these guidelines strict requirement, crack widths up to 0.010 in. (0.25 mm)
notes that a portion of the cracks in the structure might have are occasionally observed in Tx-girders currently fabricated
widths that exceed these limits. using 0.6 in. (15.2 mm) diameter strands and containing
No cracks within the end regions of the girders exceeded identical detailing for mild-steel reinforcement. Therefore,
the limit recognized by ACI 224R as tolerable for humidity the implementation of 0.7 in. (17.8 mm) diameter strands on
and soil exposure. In Tx46-I and Tx70-I, isolated cracks a 2 x 2 in. (51 x 51 mm) grid in Tx-girders does not appear
exceeding 0.007 in. (0.18 mm) were observed. However, as to cause new concerns regarding the serviceability or dura-
noted previously, a few cracks with widths greater than the bility of pretensioned girders.
listed limits are considered acceptable. Therefore, according

98 ACI Structural Journal/January 2018


Fig. 8—Stresses within end-region reinforcement of specimens. Gray shades in figure represent h/4 distance from end face.
(Note: 1 in. = 25.4 mm; 1 ksi = 6.9 MPa.)
Stresses in transverse reinforcement Tx46-II, which was designed to represent critical bursting
Figure 8 provides a graphical representation of conditions, smaller stresses were detected compared to
stresses inferred from strain measurements in the transverse Tx46-I, despite a greater prestressing force. Large stresses
reinforcement of all four specimens. Detailed information were observed only at the interface between the web and the
regarding the stresses in the transverse reinforcement is bottom flange in this specimen.
provided in Appendix D. As can be seen in Fig. 8, among In the specimens constructed with conventional strand
Tx46-I, Tx70-I, and Tx70-II, which had conventional strand patterns, the stresses in transverse reinforcement were found
patterns, the increase in the transverse reinforcement stress to diminish very quickly with distance from the end face of
levels is correlated with the increase in the total prestress the specimen. Stresses greater than 15 ksi (103 MPa) were
force. The maximum stress level in the specimens was observed only in the first three stirrups, which were located
26 ksi (179 MPa), which was observed in Tx70-II. The within 8.5 in. (220 mm) from the end face of the girder. Typi-
greatest stresses in all specimens were observed at the inter- cally, diaphragms are not used at supports for Tx-girders.
face between the web and the bottom flange, mostly due to However, in any bridge application, this distance is normally
the change in geometry and the flow of the stresses from in the overhang segment—that is, outside the main span.
the bottom flange into the web. However, the stresses also Therefore, while excessive cracking in this region might
remained large at the centroids of these three specimens. In have negative effects on the durability of the girder, these

ACI Structural Journal/January 2018 99


stresses are not expected to affect the load-carrying perfor- Table 7—Splitting resistance in specimens
mance under in-service conditions. Typical detailing used according to AASHTO LRFD4 compared to initial
for Tx-girders includes a 9 in. (230 mm) overhang segment, prestressing force
and an 8 in. (200 mm) support width for Tx46 and a 9 in. Specimen As,h/4, in.2 Pr = fsAs,h/4, kip Pi, kip Pr/Pi, %
(230 mm) support width for Tx70 girders. Using this
Tx46-I 5.12 102.4 1614 6.3
detailing, the first four stirrups in the girders are not expected
to be mobilized under external loads. The stresses in other Tx46-II 5.12 102.4 2024 5.1
stirrups were generally less than 10 ksi (69 MPa). Tx70-I 7.68 153.6 1796 8.6
Tx46-II exhibited transverse stresses that extended over Tx70-II 7.68 153.6 2738 5.6
3 ft (910 mm) into the beam. This distance was considerably
greater than that in the other three specimens. The stresses Notes: fs is taken as 20 ksi (138 MPa); 1 in.2 = 645 mm2; 1 kip = 4.4 kN.

were primarily concentrated along the web-bottom flange


interface, and only small stresses were detected at the centroid
depth and at the interface between the web and the top flange.
The maximum stresses were not observed in the first few
stirrups, but in the stirrups that were farther away from the
end face of the girder. Large stresses within this region are
known to increase the likelihood of horizontal shear distress.23
However, the magnitudes of these stresses were generally
limited to 10 ksi (69 MPa), which was small compared to
stresses observed in the other three specimens. These observa-
tions are consistent with the results obtained by O’Callaghan18
from Tx-girders that employed 0.6 in. (15.2 mm) diameter
strands on a layout resembling that of Tx46-II.
If the specimens with conventional strand patterns in
the current test program were to be fabricated using 0.5 or
0.6 in. (12.7 or 15.2 mm) diameter strands, more strands Fig. 9—Comparison between transverse forces determined
would be needed to achieve the same flexural capacity. in this study and those in database of bursting and spalling
Consequently, the distance between the centroid of strands stresses (adapted from Dunkman24 with additional data
and the web-bottom flange interface would be decreased, points). (Note: 1 kip = 4.45 kN.)
resulting in stress distribution similar to that in Tx46-II or
in girders reported by O’Callaghan.18 Such a change in the the Tx46 specimens and the first six pairs of R- and S-bars in
configuration of the girder is expected to result in a greater the Tx70 specimens. Observations from Fig. 8 support this
spread of transverse stresses and potentially large stirrup selection. In specimens with conventional strand patterns,
stresses in the main span of the girder rather than in the noticeable stresses were observed in the reinforcement
overhang region. Therefore, the use of 0.7 in. (17.8 mm) located within the h/4 region, which is shown by the gray
diameter strands might be beneficial in containing the trans- shade in the figure.
verse stresses due to prestress transfer within the overhang Figure 8 shows that maximum detected stresses in the
region, therefore preventing these end-region stresses from Tx70 specimens in this test program exceeded 20 ksi
affecting the performance of the girder under shear-critical (138 MPa). However, because the stresses diminished
loading conditions. very quickly with distance from the end face, the average
Section 5.10.10 in AASHTO LRFD4 describes require- stresses within the bars located in the first h/4 distance
ments for resisting the tensile forces generated in the end from the end face was less than 20 ksi (138 MPa); there-
regions of pretensioned elements due to bursting and fore, all girders comprising this test program effectively met
spalling stresses. In these provisions, the resistance of the the stress limit defined in AASHTO LRFD. It is important
end region against vertical tensile forces is referred to as to note that the area of mild-steel reinforcement provided
“splitting resistance”, and no distinction is made between within the h/4 distance of Tx-girders considerably exceeds
forces generated due to bursting and spalling stresses. Split- the requirements in AASHTO LRFD. As shown in Table 7,
ting resistance Pr is taken as fsAs,h/4, where fs is the stress in the use of closely spaced R- and S-bars within the end region
steel, not to exceed 20 ksi (138 MPa); As,h/4 is the total area of Tx-girders results in a splitting resistance that ranges
of reinforcement within a distance of h/4 from the end of between 5.1 and 8.6% of the initial prestressing force in
the girder; and h is the height of the girder. AASHTO LRFD each specimen.
provisions require this splitting resistance to be greater than Figure 9 includes the transverse forces developed within
4% of the total prestressing force at transfer. the distance of h/4 from the end face of the girders inves-
In this testing program, the h/4 distance is equal to 11.5 tigated in the current study and those from a database of
and 17.5 in. (292 and 445 mm) for the Tx46 and the Tx70 bursting and spalling stresses developed by Dunkman.24
specimens, respectively. On the basis of the standard details This database includes the results from eight inverted-tee
for Tx-girders, the steel required to provide the splitting specimens, four U-girders, and 53 I-girders fabricated with
resistance is limited to the first four pairs of R- and S-bars in 0.5 or 0.6 in. (12.7 or 15.2 mm) diameter strands.

100 ACI Structural Journal/January 2018


As can be seen in Fig. 9, the transverse force developed 51 mm) grid does not seem to trigger a need to modify the
over the distance of h/4 of all previous points in the data- end-region detailing in Tx-girders. Based on the guidelines
base was less than or equal to 0.04Pi, where Pi is the initial of ACI 224R-01, all specimens in this test program met the
prestress force in the bottom flange before elastic shortening criteria for use in exposure to deicing chemicals.
losses. However, in this study, noticeably greater bursting and End-region reinforcement requirements: The trans-
spalling forces were observed in all specimens except Tx46-II, verse force observed within the first h/4 length of specimens
which had a nonconventional strand pattern. In both end with conventional strand patterns exceeded 4% of the initial
regions of Tx70-I, the magnitude of bursting and spalling prestressing force. The relatively satisfactory performance
forces reached up to 5.7% of the initial prestressing force. of the girders within this test program might be attributed
to the use of considerably greater amounts of transverse
SUMMARY AND CONCLUSIONS reinforcement compared to that required by AASHTO
Four full-scale Tx-girder specimens employing 0.7 in. LRFD. The observations of this test program indicate a
(17.8 mm) diameter seven-wire strands on a 2 x 2 in. (51 x potential need to reevaluate the requirements for end-region
51 mm) grid were designed and fabricated to investigate reinforcement in AASHTO LRFD provisions when used for
their end-region behavior at the time of prestress transfer. girders employing 0.7 in. (17.8 mm) diameter strands.
The specimens were extensively instrumented to provide a A few alternatives for modifications to reinforcement
detailed picture of transverse stresses within reinforcement, detailing in Tx-girders have been examined computation-
transfer length, and prestress losses. Moreover, each spec- ally by Alirezaei Abyaneh25 to reduce the width of end-
imen was carefully examined for end-region cracking after region cracks, potentially below the strict 0.005 in. (0.13 mm)
prestress transfer and at least 28 days after prestress transfer. limit required by TxDOT. Investigating the performance of
The primary conclusions regarding the end-region behavior girders employing 0.7 in. (17.8 mm) diameter strands under
of the specimens were as follows: applied loads is beyond the scope of the current paper. The
Transfer length: The strain measurements from the shear behavior of such girders may be negatively affected by
strands revealed that the 24-hour transfer length, which diminished anchorage capacity due to greater transfer and
is known to be less than the final transfer length, slightly development lengths as well as the need to transfer a greater
exceeded the transfer lengths predicted using AASHTO horizontal shear force between the web and the bottom
LRFD and ACI 318-14 equations. While there is a possi- flange. The development of near-continuous cracks at the
bility that the presence of strain gauges contributed to an interface between the web and the bottom flange of the spec-
increased transfer length, further research on full-scale spec- imens comprising this test program could also be a source
imens is necessary to ensure accurate transfer lengths are of concern regarding the transfer of horizontal shear. These
used for the design of girders that employ 0.7 in. (17.8 mm) effects are evaluated in detail by Katz et al.3
diameter strands.
Cracking patterns: The patterns of cracks developed AUTHOR BIOS
within the specimens were similar to those observed in Jessica L. Salazar is a Bridge Designer with FIGG, Dallas, TX. She
received her BSCE from New Mexico State University, Las Cruces, NM, and
typical Tx-girders employing 0.5 or 0.6 in. (12.7 or 15.2 mm) her MSE from the University of Texas at Austin, Austin, TX. Her research
diameter strands. In all specimens, there was noticeable interests include end-region behavior of prestressed concrete elements
cracking at the web-bottom flange interface. Most cracks in and the effects of using larger-diameter strands on the performance and
economy of pretensioned concrete girders.
the test program had a width that was less than 0.007 in.
(0.18 mm), with a width of 0.008 in. (0.20 mm) observed ACI member Hossein Yousefpour is an Assistant Professor at Babol
in a few isolated locations in two specimens. Within the Noshirvani University of Technology, Babol, Iran. He received his
BSc and MSc from the University of Tehran, Tehran, Iran, and his PhD
first 28 days after prestress transfer, the length and width of from the University of Texas at Austin. His research interests include
end-region cracks increased, and new cracks appeared in all prestressed concrete, structural health monitoring, and shear strength and
girders. However, the maximum width of the cracks within time-dependent behavior of concrete structures.
each specimen did not show a noticeable change. ACI member Roya Alirezaei Abyaneh is a Structural Engineer with
Stresses within transverse reinforcement: The greatest Building Diagnostics, Inc., Houston, TX. She received her BAppS from
stresses in all specimens were observed at the interface the University of Waterloo, Waterloo, ON, Canada, and her MSE from the
University of Texas at Austin. Her research interests include prestressed
between the web and the bottom flange. A maximum stress concrete and computational simulation of reinforced concrete structures.
level of 26 ksi (179 MPa) was observed in Tx70-II. In speci-
mens with conventional strand patterns, large stirrup stresses Hyun su Kim is a PhD Student at the University of Texas at Austin. He
received his BSc and MSc from Seoul National University, Seoul, Korea, in
were observed only in the overhang region of the girders. 2008 and 2010, respectively. His research interests include shear strength
For the nonconventional strand pattern used within Tx46-II, and end-region behavior of prestressed concrete girders.
the maximum stresses were not observed in the first few stir-
Alex T. Katz is a Project Engineer with Walker Restoration Consultants,
rups but in stirrups located further away from the end face Chicago, IL. He received his BS from the University of Virginia, Char-
of the girder. lottesville, VA, and his MSE from the University of Texas at Austin. His
Necessity of modifications to detailing of end-region research interests include prestressed concrete and shear behavior of
concrete structures.
reinforcing bars: Because the typical crack widths in the
specimens did not exceed those frequently observed in ACI member Trevor D. Hrynyk is an Assistant Professor in the Depart-
girders with smaller-diameter strands, the use of 0.7 in. ment of Civil, Architectural and Environmental Engineering at the Univer-
sity of Texas at Austin. He is a member of Joint ACI-ASCE Committees 421,
(17.8 mm) diameter strands on the standard 2 x 2 in. (51 x Design of Reinforced Concrete Slabs, and 447, Finite Element Analysis of

ACI Structural Journal/January 2018 101


Reinforced Concrete Structures; and ACI Subcommittee 445-C, Punching 10. Vadivelu, J., “Impact of Larger Diameter Strands on AASHTO/PCI
Shear. His research interests include performance assessment of rein- Bulb-Tees,” master’s thesis, University of Tennessee, Knoxville, TN, 2009,
forced concrete structures, analysis and design of slabs and shells, and 114 pp.
nonlinear modeling. 11. Morcous, G.; Hanna, K.; and Tadros, M. K., “Use of 0.7-in. Diameter
Strands in Pretensioned Bridge Girders,” PCI Journal, V. 56, No. 4, 2011,
Oguzhan Bayrak, FACI, is a University Distinguished Teaching Professor pp. 65-82. doi: 10.15554/pcij.09012011.65.82
in the Department of Civil, Architectural and Environmental Engineering 12. Morcous, G.; Assad, S.; Hatami, A.; and Tadros, M. K., “Imple-
and holds the Charles Elmer Rowe Fellowship in Engineering at the mentation of 0.7 in. Diameter Strands at 2.0 x 2.0 in. Spacing in Preten-
University of Texas at Austin. He is a member of ACI Committees 341, sioned Bridge Girders,” PCI Journal, V. 59, No. 3, 2014, pp. 145-158. doi:
Earthquake-Resistant Concrete Bridges, and E803, Faculty Network Coor- 10.15554/pcij.06012014.145.158
dinating Committee; and Joint ACI-ASCE Committees 441, Reinforced 13. Texas Department of Transportation (TxDOT), “Concrete I-Girder
Concrete Columns, and 445, Shear and Torsion. Details, Standard Drawings by the Bridge Division, Austin, Texas,” Austin,
TX, 2015, 2 pp.
14. ASTM C39/C39M-14, “Standard Test Method for Compressive
ACKNOWLEDGMENTS Strength of Cylindrical Concrete Specimens,” ASTM International, West
The authors gratefully acknowledge TxDOT for financially supporting Conshohocken, PA, 2014, 7 pp.
this study, as well as Coreslab Structures in Cedar Park, TX, and Alamo 15. ASTM C469/C469M-14, “Standard Test Method for Static Modulus
Cement for donating concrete and cement that were used for the fabrication of Elasticity and Poisson’s Ratio of Concrete in Compression,” ASTM
of the specimens. Thanks are also due to R. Boehm for her assistance with International, West Conshohocken, PA, 2014, 5 pp.
the performance of the experimental program. The findings and opinions 16. ASTM A370-16, “Standard Test Methods and Definitions for
presented in this paper are those of the authors and do not necessarily reflect Mechanical Testing of Steel Products,” ASTM International, West Consho-
the views of TxDOT. hocken, PA, 2016, 49 pp.
17. ASTM A1061/A1061M-16, “Standard Test Methods for Testing
REFERENCES Multi-Wire Steel Prestressing Strand,” ASTM International, West Consho-
1. Tadros, M. K., and Morcous, G., “Impact of 0.7 Inch Diameter Strands hocken, PA, 2016, 4 pp.
on NU I-Girders,” Technical Report, Nebraska Department of Roads, 18. O’Callaghan, M. R., “Tensile Stresses in the End Regions of Preten-
Lincoln, NE, 2011, 203 pp. sioned I-Beams at Release,” master’s thesis, University of Texas at Austin,
2. Salazar, J.; Yousefpour, H.; Katz, A.; Alirezaei Abyaneh, R.; Kim, H.; Austin, TX, 2007, 265 pp.
Garber, D.; Hrynyk, T.; and Bayrak, O., “Benefits of Using 0.7 in. (18 mm) 19. Barnes, R.; Grove, J.; and Burns, N., “Experimental Assessment of
Diameter Strands in Precast, Pretensioned Girders: A Parametric Investiga- Factors Affecting Transfer Length,” ACI Structural Journal, V. 100, No. 6,
tion,” PCI Journal, V. 62, No. 6, Nov-Dec. 2017, pp. 59-75. Nov.-Dec. 2003, pp. 740-748.
3. Katz, A.; Yousefpour, H.; Kim, H.; Alirezaei Abyaneh, R.; Salazar, J.; 20. ACI Committee 224, “Control of Cracking in Concrete Structures
Hrynyk, T.; and Bayrak, O., “Shear Performance of Pretensioned Concrete (ACI 224R-01),” American Concrete Institute, Farmington Hills, MI, 2001,
I-Girders Employing 0.7 in. (17.8 mm) Strands,” ACI Structural Journal, 46 pp.
V. 114, No. 5, Sept.-Oct. 2017, pp. 1273-1284. 21. Tadros, M. K.; Badie, S. S.; and Tuan, C. Y., “Evaluation and Repair
4. American Association of State Highway and Transportation Officials Procedures for Precast/Prestressed Concrete Girders with Longitudinal
(AASHTO), “AASHTO LRFD Bridge Design Specifications,” seventh Cracking in the Web,” NCHRP Report 654, Transportation Research Board,
edition with 2015 and 2016 Interim Revisions, Washington, DC, 2016, Washington, DC, 2010, 66 pp.
2150 pp. 22. Texas Department of Transportation (TxDOT), “Standard Specifica-
5. ACI Committee 318, “Building Code Requirements for Structural tions for Construction and Maintenance of Highways, Streets, and Bridges,”
Concrete (ACI 318R-14) and Commentary (ACI 318R-14),” American Austin, TX, 2014, 919 pp.
Concrete Institute, Farmington Hills, MI, 2014, 520 pp. 23. Hovell, C.; Avedano, A.; Moore, A.; Dunkman, D.; Bayrak, O.; and
6. Russell, B. W., and Burns, N. H., “Measured Transfer Lengths of 0.5 Jirsa, J., “Structural Performance of Texas U-Beams at Prestressed Transfer
and 0.6 in. Strands in Pretensioned Concrete,” PCI Journal, V. 41, No. 5, and under Shear-Critical Loads,” Technical Report, Center for Transporta-
1996, pp. 44-65. doi: 10.15554/pcij.09011996.44.65 tion Research, Austin, TX, 2011, 355 pp.
7. Comité Euro-International Du Beton (CEB), “Anchorage Zones of 24. Dunkman, D., “Bursting and Spalling in Pretensioned U-Beams,”
Prestressed Concrete Members,” CEB Bulletin No. 181, 1987, 137 pp. master’s thesis, University of Texas at Austin, Austin, TX, 2009, 242 pp.
8. Hoyer, E., and Friedrich, E., “Beitrag zur Frage der Haftspannung in 25. Alirezaei Abyaneh, R., “Computational Modeling of Prestress
Eisenbetonbauteilen [Contribution to the Question of Bond in Concrete Transfer, End-Region Cracks, and Shear Behavior in Prestressed Concrete
Elements],” Beton und Eisen, V. 30, No. 6, 1939, pp. 107-110. I-Girders Employing Large-Diameter Strands,” master’s thesis, University
9. Dang, C. N.; Floyd, R. W.; Hale, M.; and Marti-Vargas, J. R., of Texas at Austin, Austin, TX, 2016, 95 pp.
“Measured Transfer Lengths of 0.7 in. (17.8 mm) Strands for Pretensioned
Beams,” ACI Structural Journal, V. 113, No. 3, May-June 2016, pp. 85-94.

102 ACI Structural Journal/January 2018


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 115-S08

Stepwise Bond Model Including Unconfined and Partially


Confined Hooks
by Armin Erfanian and Alaa E. Elwi

A contact interface model including hook reinforcement under fib 201010,11 allows an increase of the design bond strength
monotonic, unconfined, in-plane compression and tension is for the case of transverse pressure perpendicular to the split-
presented, which is in close agreement with experimental test ting plane. However, fib 201010,11 does not have a clear local
results. Unconfined in-plane compression accompanies unconfined bond-slip relationship for splitting failure of unconfined
hooks under the bent portion of a hook and creates splitting in the
and partially confined concrete under transverse pressure,
same plane as the hook. Although the effect of pressure on bond
whether perpendicular or parallel to the splitting plane. In
capacity of unconfined concrete is investigated in the literature,
there is little information on reliable bond-slip behavior for this addition, fib 201010,11 does not provide an ultimate value
case. fib 2010 does not provide bond-slip equations for unconfined for the splitting failure slippage of unconfined and partially
and partially confined concrete under compression, although it confined concrete under compression beyond which bond is
considers the case of confined concrete under pressure. Laborious zero. In other words, the bond-slip relationship for uncon-
effort was made in this research to formulate a procedure that fined and partially confined concrete in transverse pressure
works in the three-dimensional finite element environment without is unknown in fib 201010,11 model code and models that use
impractical limitations. In addition, a new method of computing fib 201010,11 as a basis such as the works of Zhang et al.,12 Xu
bond stress increments is proposed to be used in cases involving et al.,13 and Costa et al.14
slip-dependent normal pressure such as joints. The deficiency in fib-based models makes the modeling
Keywords: bond; confinement; in-place pressure; partial confinement; slip;
of unconfined and partially confined hooks impossible.
splitting; stepwise. Unconfined hooks are permitted by ACI 318-1415 if certain
requirements are met. Nonetheless, hooks in confined joints
INTRODUCTION meeting the requirements of ACI 318-1415 are not neces-
The present research aims at providing a research tool as sarily fully confined according to fib 201010,11 local bond
a substitution to lab experimentation. Therefore, this analyt- model provisions. In some cases, confined joints complying
ical tool should be able to simulate a range of configurations to ACI 318-1415 may meet the fib 201010,11 local bond model
regardless of whether they are currently common in prac- requirements for full confinement. In the rest of the cases,
tice. Likewise, this research tool should be able to discretize hooks in ACI-complying confined joints fall into the cate-
the effects of influencing parameters on the behavior of the gory of partially confined concrete based on fib 201010,11
structure. This paper is an effort to develop such a research local bond model. Analysis of these partially confined
tool in the field of discrete bond modeling. In contrast, the hooked joints cannot be accomplished using fib-based
smeared crack and implicit modeling assumes full compat- models. This is because fib 201010,11 does not provide a clear
ibility of steel and concrete using tension stiffening to relationship for partially confined concrete under transverse
account for bond and slip.1,2 This research uses the discrete pressure, which is in this case produced under the bend.
type of modeling. In addition, finite element models similar to the one by
One of the earliest finite element approaches with discrete Zanuy et al.,16 which are able to account for unconfined
modeling of bond has been developed by Ngo et al.,3 who transversely pressured concrete behavior, use the cohe-
modeled the bond between concrete and steel surfaces with sive zone theory instead of the bond-slip theory, which is a
spring elements. Nilson4 proposed a curve for bond in terms related but a different field of research.
of slip and used springs to connect the nodes of steel and Further, the reliability information so far published on fib
concrete elements interacting by bond. Hofstetter5 reported 2010 model10,11 does not focus on discrete bond modeling17
that Keuser et al.6 and Mehlhorn7 developed bond elements of unconfined transversely pressured concrete. Accepting
that worked better than spring elements, while Miguel et al.8 the reliability of the fib-based models does not help create a
used a combination of bond elements and spring elements finite element model for unconfined concrete with transverse
to model steel-concrete interaction. Lowes et al.9 proposed pressure based on fib 2010.10,11
a bond-slip model and an interface element for confined
concrete. fib 201010,11 presents discrete relationships for
bond and slip of unpressured unconfined and unpressured ACI Structural Journal, V. 115, No. 1, January 2018.
MS No. S-2016-328.R2, doi: 10.14359/51700949, received March 19, 2017, and
partially confined concrete as well as pressure-free and reviewed under Institute publication policies. Copyright © 2018, American Concrete
transversely pressured fully confined concrete. It is noted Institute. All rights reserved, including the making of copies unless permission is
obtained from the copyright proprietors. Pertinent discussion including author’s
that the term “confinement” is used for stirrup confinement. closure, if any, will be published ten months from this journal’s date if the discussion
is received within four months of the paper’s print publication.

ACI Structural Journal/January 2018 103


For the case of fully confined concrete, there is an abun- In this research, test results by Eligehausen et al.18 were
dance of test results and complete bond-slip relationships used for best-fitting purposes for the case of fully confined
such as the experimental results by Eligehausen et al.18 and concrete. Eligehausen et al.18 carried out a comprehen-
Shima et al.,19 who provided confinement by stirrups and sive experimental program consisting of 125 specimens of
massive covers, respectively. However, the case of uncon- confined concrete with and without transverse pressure. In
fined concrete is more complicated than the fully confined contrast to using steel tubes or large concrete covers, the
case. For the bond capacity of unconfined concrete under confinement was provided by stirrups and vertical rein-
transverse pressure, however, because of large discrepan- forcement simulating practical situations. This is the most
cies, some test results from previous researchers18,20-23 were comprehensive set of tests for confined concrete under trans-
not used herein. In addition, there are reports on the slip for verse pressure available so far. These test results have been
unconfined concrete with and without transverse pressure at referenced and used by numerous researchers in the past. A
or up to the peak load.20,22,24 These reports lack the complete recent implementation of Eligehausen et al.18 test results was
slip behavior up to failure for transversely pressured uncon- done by fib 2010,10,11 which largely adapted these test results
fined and partially confined concrete. In addition, there exist for bond-slip modeling.
numerous experiments that result in bond-slip relationships Further, the present model proposes a new method of
similar to the fib 2010 model10,11 for unconfined, trans- computing the increments of bond other than the sole use
versely unpressured, and fully confined concrete.25-29 Addi- of the slope of the bond-slip curve for the cases involving
tionally, some tests were conducted to explain the confining slip-dependent normal pressure. The procedure to imple-
stresses produced by stirrups, distribution of cracks, and ment the present model is explained along with a flowchart
crack-width-dependent bond-slip relationships for uncon- for efficient application. Experience has shown using an
fined concrete with zero transverse pressure.30,31 algorithm that updates bond stresses incrementally with
Gambarova32,33 has carried out numerous tests using respect to the current contact interface configuration is more
specimens with performed splitting cracks and a confining efficient in terms of convergence to accurate results, espe-
frame that keeps the crack opening width constant. None- cially for models including hooked bars. Thus, the proposed
theless, the confining frame in Gambarova’s32,33 experiment approach is updated in each analytical step to improve its
eliminated the possibility of failure by splitting mode, as all precision. Additionally, in the present approach, bond-slip
Gambarova’s32,33 specimens failed in pullout shearing mode. modeling of the splitting for unconfined in-plane pressure
Thus, Gambarova’s32,33 specimens cannot be considered has priority because using a proper concrete material model
unconfined concrete, as the corresponding test results do with 3-D elements eliminates the necessity to model the
not represent the splitting mode failure of transversely pres- shearing failure explicitly. An experimental test was carried
sured unconfined concrete. Instead, Gambarova’s32,33 results out in this research to help in development and optimiza-
represent the case of fully confined concrete by means of a tion and verification of the model. Further, the present model
confining frame and external transverse pressure. was validated through comparisons with other researchers’
The aforementioned discussion is important for the case experimental and analytical results.
of unconfined transversely pressured concrete in general
and particularly for the case of unconfined concrete under RESEARCH SIGNIFICANCE
hook pressure. The reason is that not all values of transverse An efficient finite element bond model is presented for
pressure fully confine splitting. Specifically, the normal discrete modeling of bond at the contact interface of steel
pressure developed under a hook parallel to the splitting and concrete for monotonic loading. The proposed model
plane does not contain the splitting failure, no matter how produces accurate results in different geometrical and
large it becomes. Hence, because of these reasons, a bond- confinement conditions. In contrast to other models, the
slip relationship for the splitting failure of unconfined trans- present model considers unconfined in-plane pressure and
versely pressured concrete cannot be created according to tension for hooks in a stepwise analysis. Using the present
Gambarova’s32,33 test results. model along with 3-D finite element package enables
Therefore, a finite element model for the analysis of researchers to simulate controlled test conditions to study
transversely pressured unconfined and partially confined the behavior of reinforcing bars without building the actual
concrete cannot be constructed solely based on previous specimens in a laboratory, which helps save a considerable
experiments. This is especially true for the case of uncon- amount of time and resources.
fined hooks, where the pressure under the bend is parallel to
the splitting plane. In this case, either of the terms “in-plane EXPERIMENTAL TEST
pressure” or “in-plane stress” is a more suitable term and is To investigate the behavior of unconfined concrete under
suggested herein. Accordingly, this research presents a finite- in-plane hook pressure, the specimen of Fig. 1 was tested in
element-based bond model that accounts for bond capacity a universal servo-controlled machine. The loading method
and slip in transversely pressured unconfined and partially was displacement control with an approximate rate of
confined concrete. In addition, the present model considers loading of 0.001 mm/s (3.94 × 10–5 in./s). Because a joint
the effect of in-plane tension, varying slip-dependent normal is a disturbed region, the strut-and-tie method was used to
pressure on the bars and hooks, and splitting of concrete for design it. Weldable, low-alloy, Type 400W bars conforming
both pressured and unpressured conditions in a three-dimen- to CSA G30.18-M92 with the nominal yield stress, fy of
sional (3-D) finite element environment. 400 MPa (58 ksi) were used. The actual yield stress fya, based

104 ACI Structural Journal/January 2018


Fig. 1—Present experimental test specimen. (Note: Dimensions are in mm; 1 mm = 0.0394 in.)
on coupon tests, was 450 MPa (65.25 ksi). The compressive
strength of concrete, fc′, was 35 MPa (5.07 ksi). The diam-
eters of the tension longitudinal and shear reinforcement
were 35 and 15 mm (1.38 and 0.6 in.), respectively. There
were two layers of 20 mm (0.78 in.)-diameter longitudinal
compression bars. The tension longitudinal reinforcement is
bent by 90 degrees in the middle with a diameter of 254 mm
(10 in.) and is hooked by 180 degrees at both ends with the
same diameter as the middle bend.
The present specimen was designed for a capacity up to
yielding of bars at the face of the joint. However, a hook
development length of approximately 60% of the full length
Fig. 2—Failed experimental test specimen.
was provided to ensure bars cannot fully develop yield
strength and, therefore, they have to fail by the loss of bond. load-deflection and stress diagrams from this test are used
Thus, a study of the effect of unconfined in-plane pressure to develop bond-confinement and bond-slip relationships.
on bond capacity becomes possible because other modes These diagrams are also used to make comparisons of the
of failure have been eliminated. Unconfined in-plane pres- developed model with the present experiment.
sure is single-sided normal pressure produced in the bend
region, resulting in splitting in unconfined concrete in the INTERFACE BOND MODEL
same plane as the hook. Thus, at the time, one face of steel is Assuming the local bond stress can be presented as a func-
under normal pressure the other is not. tion of slip and an externally applied constant normal pres-
The failure, shown in Fig. 2, was by splitting of the concrete sure, the i-th increment of the bond stress would be equal
at the joint region at the peak load and, as a result, the side to the product of a parameter iΓs and the increment of slip,
covers spalled off and then a descending branch occurred where iΓs would be the slope of the bond-slip relationship.
on load versus loading end displacement diagram in Fig. 3. However, in a more practical case of a hook pullout test, the
The strains were measured by a set of strain gauges placed normal pressure is applied internally as a consequence of slip-
on the bars from the face of the joint to the beginning of the page and depends on the slip. In this case, using the product
bend. Figures 4 and 5 show the values of average strains at of iΓs and the slip increment to compute the bond stress incre-
the beginning of the bend and at the face of the joint. The ment is not a complete solution and renders erroneous results.

ACI Structural Journal/January 2018 105


normal stress is in tension, the normal stress parameter would
be equal to zero, as stated in the following equation

 σl
 σl ≥ 0
n =  f c′ (2)
0 σl ≤ 0

Likewise, the bond stress is dependent on the state of the
two contacting surfaces. However, the bond stress depends
on the magnitude of the local normal stress as well. Tensile
normal stress exceeding the interface chemical adhesion
results in zero bond stress. This is because the contacting
surfaces are separated. Further, if the local normal tensile
Fig. 3—Comparison of present model and with present test stress does not exceed the chemical adhesion capacity, the
results. bond stress equals to the accumulation of all bond stress
increments up to the current state of analysis. The following
equation expresses this concept

Σ i τ σl ≥ −m
τ= (3)
0 σl ≤ −m

where μ is the normal chemical adhesion stress capacity.


Chemical adhesion stress parallel to the bars can be found in
the literature.34 However, in the present research, the chem-
ical adhesion stress is used in the direction perpendicular to
the steel-concrete interface. The present treatment of tension
Fig. 4—Present model versus present test for stresses at
helps maintain the stability of some cases of analysis with
beginning of bend.
in-plane tension in bent bars. In-plane tension can occur
in bent bar conditions when the contact between steel and
concrete is lost on the outside face of a hook or inside face
of a hook tail near the failure load, or when displacements
become large and the contacting surfaces start to separate.
To distinguish between full, partial, and no confinement, a
new measure Λ is introduced. In cases of full and no confine-
ment, Λ would be 1 and 0, respectively. Thus, any values
between 0 and 1 represent a partially confined case. There-
fore, as the volumetric ratio of confining reinforcement
ρ becomes larger, the parameter Λ gets closer to 1. Like-
wise, reducing the confinement volumetric ratio results in
Fig. 5—Present model versus present test for stresses at face Λ getting closer to 0. Considering f(ρ) as a function of ρ, an
of joint. expression such as (1 – e–f(ρ)) would be asymptotic to 1 and
0 for large and small values of f(ρ), respectively. Therefore,
Therefore, to have a mathematically correct solution, a new the following equation applies
additional parameter iΓn is introduced in this research, which
is associated to the changes in slip-dependent normal stress. Λ = 1 – e–f(ρ) (4)
Equation (1) presents the preceding discussion
f(ρ) = 10(ρ ·ρs–1)1.5 (5)
τ = iΓs · is + iΓn · in (1)
i

where iτ and is denote the incremental values of the bond


stress τ and slip s. The normal stress parameter in is equal Having the values of bond capacities for unconfined
to the increment of the normal stress parameter n, which is concrete under normal pressure, τupc, and confined concrete
defined as the ratio of the of normal stress to the compressive under normal pressure, τcpc, the bond capacity for partially
strength of concrete, fc′. The normal stress parameter depends confined concrete under normal pressure is obtained as follows
on the steel-concrete contact surfaces. When the local normal
stress σl is compressive or zero, the normal stress parameter τec = τcpcΛ + τupc(1 – Λ) (6)
is computed as the stated ratio. In contrast, when the local

106 ACI Structural Journal/January 2018


As a comparison, in the fib 2010 model,10,11 computation
of the bond capacity for partially confined concrete under
normal pressure is not provided. Considering the concept
of Eq. (1) and (6), Fig. 6 illustrates the advantage of using
Eq. (1), proposed in this research, for the computation of the
bond stress incrementally. It is clear in Fig. 6 that the imple-
mentation of Eq. (1) is more precise than the conventional
application of the total values of slip and normal pressure
into a bond-slip relationship. For simplicity of comparison, in
Fig. 6, a hypothetical bilinear bond-slip relationship is shown.
It is expected that in every step of the analysis at different
locations, different normal pressures and, consequently,
different bond capacities, exist. The proposed bond capacity
for unconfined concrete under normal pressure is given by
Eq. (7), which is obtained by a large number of trials and
optimization of the parameters on the finite element model
of the present specimen bent bar region, which is stirrup-free
but develops unconfined in-plane normal pressure. Fig. 6—Incremental versus total computation of bond with
hypothetical bilinear bond-slip relation.
τupc = τuc(α1n3 + α2n2 + α3n + 1) (7)
η = 2.4τucΛ(3β1n2 + 2β2n + β3) + τuc(1 – Λ)(3α1n2 + 2α2n + α3)
where τuc is the unconfined bond capacity. Parameters α1, α2, (11)
and α3 are optimized as 0.4086, –2.4855, and 3.4769, respec-
tively. The unconfined bond capacity τuc, which is dependent where
on the cover to bar diameter ratio and the tensile strength
of concrete, can be obtained from the literature.36 Equations τ ec
sm = (12)
(1) and (7) make hook and bent bar analysis possible for κ
unconfined concrete where there is normal pressure parallel
to the splitting failure at the inside face of the bent bar. This sm2 = c1(sm – sc) + sc (13)
task cannot be accomplished with the fib 2010 model,10,11 as
it does not provide an equation to account for τupc, nor does su = [c2(sm – sc) + sc](1 – Λ)–1 (14)
it provide a relationship of bond with slip for unconfined
concrete under normal pressure. In the aforementioned equations, splitting of uncon-
To model the behavior of confined concrete under pres- fined and partially confined concrete under pressure are
sure, Eq. (8) is proposed by the authors best-fitting the modeled explicitly, and su is the amount of slip beyond
experimental test results by Eligehausen et al.18 which bond is zero. For fully confined concrete, su generates
a very large number before which the shearing failure will
τcpc = 2.4τuc(β1n3 + β2n2 + β3n + 1) (8) happen. However, there is no need to model the shearing
failure in the present approach because the shearing mode
where parameters β1, β2, and β3 are best-fitted to 2.98, –3.75, is not a bond failure, but a concrete failure. In other words,
and 1.61, respectively. It is obvious that n has a limit, but using an appropriate material model for concrete, when the
there is no need to enforce an explicit limit to Eq. (7) and concrete at the interface reaches its critical value of shearing
(8). Since using an acceptable material model, the concrete deformation, it fails even without explicit modeling of the
under the hook will crush under extremely large values of shearing mode.
normal pressure. As the result of parameter optimization, κ is a constant that
Here, parameters iΓs and iΓn are defined is calibrated as 250 N/mm3 (9.2 × 105 lbf/in.3) to produce an
ascending branch as close as possible to the load-deflection
κ s ≤ sm diagram of Fig. 3. This value results in acceptable analytical
0 responses for normal-strength concrete. The constants c1, c2,
 sm < s ≤ sm 2
and sc are obtained as 1.468, 21.11, and 0.018 mm (7.09 ×
i
Γ s =  τ ec (9) 10–4 in.), respectively, by calibration of the parameters over
 s − s (1 − Λ ) sm 2 < s ≤ su
a large number of analyses of the presented specimen.
 m2 u
0 s > su Figure 7 presents the flowchart for the proposed procedure
implemented to each analysis step. In this flowchart, the
area associated with each node is specified first. Then, the
0 s ≤ sm orientation of the local frame of reference is given using the
η sm < s ≤ sm 2 direction cosines of the tangent to the bar. The increments of
 normal pressure and slip are computed next.
i
Γn =  η
 η + s − s ( s − sm 2 )(1 − Λ ) sm 2 < s ≤ su (10)
 m2 u

0 s > su
ACI Structural Journal/January 2018 107
Fig. 7—Flowchart.
The normal pressure is considered proportional to the Furthermore, explicit dynamic analysis in quasi-static form
amount of penetration by the steel surface into the concrete was carried out using the software ABAQUS.42 In all cases,
surface. If the reference coordinate system for the inter- 3-D solid elements were implemented to model concrete and
face is different from the local coordinate system, which is main tension steel while embedded truss elements were used
usually the case, a conversion needs to be made to obtain the to model stirrups and compression bars. In addition, tension
slip and its increment in the local frame of reference using steel and concrete are connected through a contact interface.
the direction cosines of the interface reference frame and the Further, the present interface bond model was programmed
local reference system assigned at the beginning of the flow- into the software as a user subroutine. Nonetheless, the
chart. A proper contact algorithm should be used wherein damaged plasticity material model43,44 was implemented
the interface coordinate system is updated with the contact for concrete. In contrast to the bond action, the formation
surface deformations. of cracks was not modeled discretely. In other words, cracks
The next step is to calculate the bond capacity based on were indirectly considered through the concrete material
the values obtained in preceding steps. Next is to compute model. To eliminate the analysis sensitivity to the mesh
the increment of bond stress depending on the increments of size, the fracture energy was divided by the characteristic
slip and normal stress. The last stage is to store the current length.43,45 Variable mass scaling42 was implemented to
values of bond in the local coordinate system for every loca- limit the stress wave transit of a time step to less than the
tion, as they will be used in the next step. smallest element dimension. The stable time steps used in
this procedure are shown in Table 2 along with other model
VALIDATION OF MODEL configurations. The increments of the prescribed displace-
The presented model and algorithm was used to analyze ment are generally small, as shown in Table 2. This type of
different specimens with straight and bent bars from present analysis is not iterative. Additionally, symmetry of speci-
and past research. Experimental specimens from the present mens is accounted for by modeling half of the specimens
researchers Untrauer and Henry,22 Lormanometee,20 Elige- with respect to each plane of symmetry and applying the
hausen et al.18 and Minor and Jirsa37 were considered for appropriate boundary conditions on the planes of symmetry.
validation. In addition, analytical models from Luccioni et For example, Fig. 8 shows the finite element mesh used to
al.,38 Lowes et al.,9 Desir et al.,39 Lettow,40 and Mendes and model the present specimen.
Castro41 were used for comparison. Table 1 shows the cases In this research, priority is given to the modeling of
studied as well as the comparison of test and analysis results. unconfined concrete and, thus, most of the verification cases

108 ACI Structural Journal/January 2018


Table 1—Specimens modeled
Case Specimen Type db/dba fc′/fca′ fy/fya n0 Pmax-Test, kN (kip) Pmax-FE, kN (kip) Pmax-FE/Pmax-Test
1 Present specimen J 1.40 1.17 1.00 0.00 970.00 (211.50) 921.00 (207.00) 0.95
2 36625 22
SP 0.75 0.83 1.59 0.25 118.26 (26.61) 111.89 (25.18) 0.95
3 46600 22
SP 0.75 1.07 1.59 0.00 81.87 (18.42) 80.05 (18.01) 0.98
4 4762022 SP 0.75 1.10 1.59 0.20 148.28 (33.36) 127.36 (28.65) 0.86
5 6-6-52-0-A420 SP 0.75 1.19 1.00 0.00 80.05 (18.01) 87.33 (19.65) 1.09
6 6-6-52-48-A2 20
SP 0.75 1.19 1.00 0.48 172.84 (38.89) 179.21 (40.32) 1.04
7 6-6-52-24-A6 20
SP 0.75 1.19 1.00 0.24 171.93 (38.68) 154.65 (34.80) 0.90
8 6-6-39-0-D120 SP 0.75 0.90 1.00 0.00 73.68 (16.58) 71.86 (16.17) 0.98
9 9-6-41-0-G120 SP 1.30 0.95 1.00 0.00 85.19 (19.17) 84.23 (18.95) 0.99
10 9-6-48-43-J6 20
SP 1.30 1.10 1.00 0.44 255.58 (57.51) 212.98 (47.92) 0.83
11 1.4.2 9,18,38,39
SP 1.00 1.00 1.33 0.00 60.80 (13.68) 55.74 (12.54) 0.92
12 6.49,18,40,41
SP 1.00 1.00 1.33 0.45 172.28 (38.76) 172.28 (38.76) 1.00
13 7-8.5-90-1.5 A37 HP 0.88 1.27 1.00 0.00 155.25 (34.90) 150.59 (34.98) 0.97
14 7-8.5-90-1.5 B 37
HP 0.88 1.33 1.00 0.00 155.25 (34.90) 155.25 (34.90) 1.00
15 7-8.5-90-2.0 A 37
HP 0.88 0.87 1.00 0.00 168.45 (37.87) 178.56 (40.14) 1.06
16 7-8.5-90-2.0 B 37
HP 0.88 1.08 1.00 0.00 181.84 (40.88) 198.20 (44.55) 1.09
17 5-4.5-90-1.5 A37 HP 0.63 1.24 1.00 0.00 86.29 (19.40) 80.25 (18.04) 0.93
18 9-8.3-90-3.0 A 37
HP 1.13 0.76 1.00 0.00 236.24 (53.11) 229.16 (51.52) 0.97
19 9-8.3-90-4.5 A 37
HP 1.13 0.62 1.00 0.00 191.50 (43.05) 177.15 (39.82) 0.93
20 Present Model HP 1.18 1.00 1.00 0.00 282.74 (63.56 )
* *
279.91 (62.93) 0.99
*
Capacity from ACI 318-14.
Notes: db is bar diameter; Pmax-Test is maximum load from test; Pmax-FE is maximum load from analysis; db is bar diameter; dba = 25.4 mm (1 in.); fca′ = 30 MPa (4.35 ksi); fya = 400 MPa
(58 ksi); n0 is external pressure divided by fc′; J is joint; SP is straight bar pullout; HP is hook pullout.

involve unconfined concrete. Therefore, except for Elige- material properties at large slip values associated with the
hausen’s Specimen 6.418 and the straight arms of the present pullout shearing failure. Overall, the comparisons in the
specimen outside the joint, which are confined, the rest of aforementioned tables and figures show that the present
the specimens are unconfined. In the analysis, the nominal model renders results with close agreement to the present
yield stress of steel and compressive strength of concrete and past experiments. Figure 16 shows an unconfined hook
given by the corresponding researcher is used. pullout simulation created with the present model compared
Figures 3 to 5 and Fig. 9 to 15 illustrate the stated compar- with the hook capacity from ACI 318-14.15 For the model in
isons. Figure 14 compares the present model results with Fig. 16, the required hook development length for yielding of
experimental and analytical results from other researchers. steel is provided according to ACI 318-14,15 with two hooks
This figure presents the analysis for unconfined concrete spaced at 280 mm (11 in.) and top, side, and hook tail back
without transverse pressure. It is evident from Fig. 14 that clear covers of 68, 45, and 100 mm (2.68, 1.77, and 3.94 in.),
the present model results are in close agreement with the respectively. Bar size and material properties for the model
experimental results. A few other presented cases of compar- of Fig. 16 are given in Table 1, as load and strength reduction
ison include unconfined concrete under transverse pressure factors are set to unity and the boundary conditions were
including unconfined hooks. These analysis cases were applied in the same manner as Minor and Jirsa37 enforced
compared to experimental results. Currently the authors do them. It is clear from Fig. 16 that the predicted capacity from
not know of any other bond-slip model capable of modeling the present model and ACI 318-1415 are in close agreement.
unconfined concrete under transverse pressure, in-plane
and out-of-plane. In Fig. 15, the present model exhibits a SUMMARY AND CONCLUSIONS
softer behavior than the experiment at large slip values. A stepwise constitutive model for the bar-concrete inter-
This is because the shearing failure in the present model face is proposed THAT accounts for all influential param-
is dependent on the material properties. In other words, an eters, including bond-slip for unconfined and partially
improvement is expected if the exact material curves from confined in-plane pressure and tension. The model is
the test are applied to the analysis. However, some of the suggested for 3-D finite element step-by-step analysis to
material properties including the stress-fracture energy enable researchers to simulate laboratory experiment results
curve have been approximated according to the literature.45 without expending the associated time and resources. The
This is because of the unavailability of such data from the proposed equations are to be used in conjunction with a
test results. Therefore, the present approach is sensitive to contact algorithm. An efficient procedure on how to imple-

ACI Structural Journal/January 2018 109


Table 2—Configurations of specimens
Δδp × 104,
CN CET Nec Nnc Dnc SET Nes Nns Dns Δts × 104, s Δt × 104, s t, s Ns × 10–2 mm (in.) δp, mm (in.)
1 B/W 3062/888 8/6 3/3 B/T 82/172 8/2 3/3 5.0 3.0 10.0 200 4.20 (0.17) 14.00 (0.55)
2 W 372 6 3 B 12 8 3 5.0 3.0 10.0 200 0.38 (0.01) 1.25 (0.049)
3 W 372 6 3 B 12 8 3 5.0 3.0 10.0 200 0.18 (0.01) 0.60 (0.024)
4 W 372 6 3 B 12 8 3 5.0 3.0 10.0 200 0.27 (0.01) 0.91 (0.036)
5 W 372 6 3 B 12 8 3 50.0 5.0 10.0 200 0.48 (0.02) 0.95 (0.038)
6 W 372 6 3 B 12 8 3 50.0 5.0 10.0 200 1.00 (0.04) 2.00 (0.078)
7 W 372 6 3 B 12 8 3 50.0 5.0 10.0 200 1.00 (0.04) 2.00 (0.078)
8 W 372 6 3 B 12 8 3 50.0 5.0 10.0 200 0.35 (0.01) 0.69 (0.027)
9 W 372 6 3 B 12 8 3 50.0 5.0 10.0 200 0.30 (0.01) 0.60 (0.024)
10 W 372 6 3 B 12 8 3 50.0 5.0 10.0 200 0.48 (0.02) 0.96 (0.038)
11 W 6200 6 3 W 40 6 3 10.0 3.0 20.0 200 0.60 (0.02) 4.00 (0.16)
12 W 6200 6 3 W/T 40/76 6/2 3/3 10.0 3.0 20.0 200 1.38 (0.05) 9.16 (0.36)
13 W 5130 6 3 W 38 6 3 50.0 50.0 5.0 10 8.50 (0.33) 0.85 (0.33)
14 W 5149 6 3 W 38 6 3 50.0 50.0 5.0 10 10.00 (0.39) 1.00 (0.039)
15 W 5149 6 3 W 38 6 3 50.0 50.0 5.0 10 13.70 (0.54) 1.37 (0.054)
16 W 5149 6 3 W 38 6 3 50.0 50.0 5.0 10 12.20 (0.48) 1.22 (0.048)
17 W 5149 6 3 W 38 6 3 50.0 50.0 5.0 10 14.70 (0.58) 1.47 (0.058)
18 W 5149 6 3 W 38 6 3 50.0 50.0 5.0 10 13.70 (0.54) 1.37 (0.054)
19 W 5149 6 3 W 38 6 3 50.0 50.0 5.0 10 10.90 (0.43) 1.09 (0.043)
20 W 7332 6 3 W 60 6 3 20.0 4.0 5.0 25 2.80 (0.11) 3.50 (0.14)

Notes: CN is case number; CET is concrete element type; Nec is number of concrete elements; Nnc is number of nodes per concrete element; Dnc is degrees of freedom per concrete
node; SET is steel element type; Nes is number of steel elements; Nns is number of nodes per steel element; Dns is degrees of freedom per steel node; Δts is stable time step; Δt
is time increment; ts is total time period; Ns is number of steps; Δδp is increment of prescribed displacement; δp is total prescribed displacement; B is brick element; W is wedge
element; T is truss element.

Fig. 9—Present finite element approach versus


Lormanometee’s20 test results.
ment between the presented approach and test results. A
flowchart of the present procedure is also demonstrated.
Fig. 8—Finite element mesh of present specimen. A sequel to this research applies the present model to a
parametric study where the response of bars in specific
ment the relationships is presented. Although the bond-slip bent and straight configurations independent of stirrups is
parameters were optimized for the present experiment, it is studied along with the assessment of the corresponding ACI
shown that they work well for other cases without modifi- 318-1415 provisions and the recommendation of appropriate
cation. The present procedure leads to accurate results for modifications. It is imperative in a parametric study to assess
different types of reinforcement configurations. Compari- the influence of each parameter on the behavior of structure
sons have been made between this work and experiments independently of other parameters, and the present bond
from the literature. These comparisons render close agree- model provides a means to that purpose. In other words, this

110 ACI Structural Journal/January 2018


Fig. 10—Present finite element approach versus
Lormanometee’s20 test results.
Fig. 14—Comparison with Eligehausen’s test18 for uncon-
fined concrete.

Fig. 11—Present finite element approach versus


Lormanometee’s20 test results.

Fig. 15—Comparison with Eligehausen’s test18 for fully


confined concrete.

Fig. 12—Present finite element approach versus Minor’s37


test results.
Fig. 16—Comparison of the present hook pullout model with
ACI 318-14.15
parametric study could not be completed without the imple-
mentation of the present model.

AUTHOR BIOS
Armin Erfanian is a Senior Structural and R&D Engineer at AETECH
Consulting, Calgary, AB, Canada. He received his BS from Amirkabir
University of Technology, Tehran, Iran; his MS from Ferdowsi University
of Mashhad, Mashhad, Iran; and his PhD from the University of Alberta,
Edmonton, AB, Canada. His research interests include anchorage of bars
in reinforced concrete and advanced finite element modeling and analysis.

Alaa E. Elwi is a Structural Engineering Professor Emeritus at the Univer-


sity of Alberta. He received his BS from the University of Cairo, Cairo,
Egypt; and his MS and PhD from the University of Alberta. His research
Fig. 13—Present finite element approach versus Minor’s37 interests include bond in reinforced concrete, composite structures,
test results. dynamics of structures, and advanced finite element method.

ACI Structural Journal/January 2018 111


ACKNOWLEDGMENTS 22. Untrauer, R., “E., and Henry, R. L., “Influence of Normal Pressure on
The authors wish to express their gratitude and sincere appreciation to the Bond Strength,” ACI Journal Proceedings, V. 62, No. 5, May 1965, pp. 577-586.
Natural Sciences and Engineering Research Council of Canada and Cement 23. Robins, P. J., and Standish, I. G., “The Influence of Lateral Pressure
Association of Canada for funding this research. upon Anchorage Bond,” Magazine of Concrete Research, V. 36, No. 129,
1984, pp. 195-202. doi: 10.1680/macr.1984.36.129.195
24. Rucka, M., and Wilde, K., “Experimental Study on Ultrasonic
REFERENCES Monitoring of Splitting Failure in Reinforced Concrete,” Journal of
1. Link, R. A.; Elwi, A. E.; and Scanlon, A., “Biaxial Tension Stiffening Nondestructive Evaluation, V. 32, No. 4, 2013, pp. 372-383. doi: 10.1007/
Due to Generally Oriented Reinforcing Layers,” Journal of Engineering s10921-013-0191-y
Mechanics, ASCE, V. 115, No. 8, 1989, pp. 1647-1662. doi: 10.1061/ 25. Solomos, G., and Berra, M., “Rebar Pullout Testing under Dynamic
(ASCE)0733-9399(1989)115:8(1647) Hopkinson Bar Induced Impulsive Loading,” Materials and Structures,
2. He, X. G., and Kwan, A. K. H., “Modeling Dowel Action of Rein- V. 43, No. 1-2, 2010, pp. 247-260. doi: 10.1617/s11527-009-9485-z
forcement Bars for Finite Element Analysis of Concrete Structures,” 26. Ghandehari, M.; Krishnaswamy, S.; and Shah, S., “Bond-Induced
Computers & Structures, V. 79, No. 6, 2001, pp. 595-604. doi: 10.1016/ Longitudinal Fracture in Reinforced Concrete,” Journal of Applied
S0045-7949(00)00158-9 Mechanics, V. 67, No. 4, 2000, pp. 740-748. doi: 10.1115/1.1313822
3. Ngo, D., and Scordelis, A. C., “Finite Element Analysis of Reinforced 27. Murcia-Delso, J., and Shing, P., “Bond-Slip Model for Detailed
Concrete Beams,” ACI Journal Proceedings, V. 64, No. 3, Mar. 1967, Finite-Element Analysis of Reinforced Concrete Structures,” Journal of
pp. 152-163. Structural Engineering, ASCE, V. 141, No. 4, 2015, p. 04014125. doi:
4. Nilson, A. H., “Nonlinear Analysis of Reinforced Concrete by the 10.1061/(ASCE)ST.1943-541X.0001070
Finite Element Method,” ACI Journal Proceedings, V. 65, No. 9, Sept. 28. Desnerck, P.; Lees, J. M.; and Morley, C. T., “Bond Behaviour of
1968, pp. 757-766. Reinforcing Bars in Cracked Concrete,” Construction and Building Mate-
5. Hofstetter, G., and Mang, H. A., Computational Mechanics of Rein- rials, V. 94, 2015, pp. 126-136. doi: 10.1016/j.conbuildmat.2015.06.043
forced Concrete Structures, Vieweg & Sohn Verlagsgesellschaft mbH, 29. Lahnert, B. J.; Houde, J.; and Gerstle, K. H., “Measurement of Slip
Wiesbaden, Germany, 1995, 366 pp. Between Steel and Concrete Core,” ACI Journal Proceedings, V. 83, No. 6,
6. Keuser, M., and Mehlhorn, G., “Finite Element Models for Bond May-June 1986, pp. 974-982.
Problems,” Journal of Structural Engineering, ASCE, V. 113, No. 10, 1987, 30. Maeda, M.; Otani, S.; and Aoyama, H., “Effect of Confinement
pp. 2160-2173. doi: 10.1061/(ASCE)0733-9445(1987)113:10(2160) on Bond Splitting Behavior in Reinforced Concrete Beams,” Struc-
7. Mehlohrn, G., “Some Developments for Finite Element Analysis of tural Engineering International, V. 5, No. 3, 1995, pp. 166-171. doi:
Reinforced Concrete Structures,” Proceedings of the Second International 10.2749/101686695780601042
Conference on Computer Aided Analysis and Design of Concrete Struc- 31. Lura, P.; Plizzari, G.; and Riva, P., “Splitting Crack Propagation
tures, Pineridge Press, Swansea, UK, 1990, pp. 1319-1336. in Pull-out Tests,” Proceedings of FramCos IV, Fracture Mechanics of
8. Miguel, P. F.; Jawad, M. A.; and Fernandez, M. A., “A Discrete- Concrete Structures, R. de Borst, J. Mazars, and J. Pijaudier-Cabot, eds.,
Crack Model for the Analysis of Concrete Structures,” Proceedings of the Balkema, Paris, France, 2001, pp. 999-1006.
Second International Conference on Computer Aided Analysis and Design 32. Gambarova, P.; Rosati, G. P.; and Zasso, B., “Steel-to-Concrete Bond
of Concrete Structures, Pineridge Press, Swansea, UK, 1990, pp. 847-908. after Concrete Splitting: Test Results,” Materials and Structures, V. 22,
9. Lowes, L. N.; Moehle, J. P.; and Govindjee, S., “Concrete-Steel Bond No. 1, 1989, pp. 35-47. doi: 10.1007/BF02472693
Model for Use in Finite Element Modeling of Reinforced Concrete Struc- 33. Gambarova, P.; Rosati, G. P.; and Zasso, B., “Steel-to-Concrete Bond
tures,” ACI Structural Journal, V. 101, No. 4, July-Aug. 2004, pp. 501-511. after Concrete Splitting: Constitutive Laws and Interface Deterioration,”
10. “fib Model Code for Concrete Structures 2010,” Ernst & Sohn, Materials and Structures, V. 22, No. 5, 1989, pp. 347-356. doi: 10.1007/
Berlin, Germany, 2013. 500 pp. BF02472505
11. Cairns, J., “Bond and Anchorage of Embedded Steel Reinforcement 34. Feldman, L. R., and Bartlett, M., “Bond Strength Variability in
in the fib Model Code 2010,” Structural Concrete, V. 16, No. 1, 2015, Pullout Specimens with Plain Reinforcement,” ACI Structural Journal,
pp. 45-55. doi: 10.1002/suco.201400043 V. 102, No. 6, Nov.-Dec. 2005, pp. 860-867.
12. Zhang, X.; Wang, L.; Zhang, J.; and Liu, Y., “Model for Flexural 35. ACI Committee 318, “Building Code Requirements for Structural
Strength Calculation of Corroded RC Beams Considering Bond-Slip Concrete (ACI 318-11) and Commentary,” American Concrete Institute,
Behavior,” Journal of Engineering Mechanics, ASCE, V. 142, No. 7, 2016, Farmington Hills, MI, 2011, 503 pp.
pp. 1-11. doi: 10.1061/(ASCE)EM.1943-7889.0001079 36. Tepfers, R., “Cracking of Concrete Cover along Anchored Deformed
13. Xu, L.; Hai, T.; and King, L. C., “Bond Stress-Slip Prediction Reinforcing Bars,” Magazine of Concrete Research, V. 31, No. 106, 1979,
under Pullout and Dowel Action in Reinforced Concrete Joints,” ACI pp. 3-12. doi: 10.1680/macr.1979.31.106.3
Structural Journal, V. 111, No. 4, July-Aug. 2014, pp. 977-988. doi: 37. Minor, J., and Jirsa, J., “Behavior of Bent Bar Anchorages,” ACI
10.14359/51686816 Structural Journal, V. 72, No. 4, Apr. 1975, pp. 141-149.
14. Costa, R.; Providência, P.; and Dias, A., “Anchorage Models for 38. Luccioni, B. M.; López, D. E.; and Danesi, R. F., “Bond-
Reinforced Concrete Beam-Column Joints under Quasi-Static Loading,” Slip in Reinforced Concrete Elements,” Journal of Structural Engi-
ACI Structural Journal, V. 113, No. 3, May-June 2016, pp. 503-514. doi: neering, ASCE, V. 131, No. 11, 2005, pp. 1690-1698. doi: 10.1061/
10.14359/51688759 (ASCE)0733-9445(2005)131:11(1690)
15. ACI Committee 318, “Building Code Requirements for Structural 39. Désir, J.-M.; Romdhane, M. R. B.; Ulm, F.-J.; and Fairbairn,
Concrete (ACI 318-14) and Commentary (ACI 318R-14),” American E. M. R., “Steel-Concrete Interface: Revisiting Constitutive and Numerical
Concrete Institute, Farmington Hills, MI, 2014, 520 pp. Modeling,” Computers & Structures, V. 71, No. 5, 1999, pp. 489-503. doi:
16. Zanuy, C.; Curbach, E. H. M.; and Lindorf, A., “Finite Element 10.1016/S0045-7949(98)00308-3
Study of Bond Strength between Concrete and Reinforcement under 40. Lettow, S., “Ein Verbundelement für nichtlineare Finite Elemente
Uneven Confinement Condition,” Structural Concrete, V. 14, No. 3, 2013, Analysen - Anwendung auf Übergreifungsstöße. Dissertation,” Institut für
pp. 260-270. doi: 10.1002/suco.201200019 Werkstoffe im Bauwesen, Universität Stuttgart, Stuttgart, Germany, 2006,
17. Cervenka, V., “Reliability-Based Non-Linear Analysis According to 196 pp.
fib Model Code 2010,” Structural Concrete, V. 14, No. 1, 2013, pp. 19-28. 41. Mendes, L. A. M., and Castro, M. S. S., “A New RC Bond Model
doi: 10.1002/suco.201200022 Suitable for Three-Dimensional Cyclic Analyses,” Computers & Struc-
18. Eligehausen, R.; Popov, E. P.; and Bertero, V. V., “Local Bond Stress- tures, V. 120, 2013, pp. 47-64. doi: 10.1016/j.compstruc.2013.01.007
Slip Relationships of Deformed Bars under Generalized Excitations,” 42. ABAQUS, “Abaqus/CAE User’s Manual.” Dassault Systémes, 2012.
Report No. UCB/EERC-83/23, Earthquake Engineering Research Center, 43. Oller, S.; Onate, E.; Oliver, J.; and Lubliner, J., “Finite Element
University of California, Berkeley, Berkeley, CA, 1983, 169 pp. Nonlinear Analysis of Concrete Structures Using a Plastic-Damage Model,”
19. Shima, H.; Chou, L. L.; and Okamura, H., “Micro and Macro Models Engineering Fracture Mechanics, V. 35, No. 1-3, 1990, pp. 219-231. doi:
for Bond in Reinforced Concrete,” Journal of the Faculty of Engineering, 10.1016/0013-7944(90)90200-Z
University of Tokyo, Tokyo, Japan, V. 39, No. 2, 1987, pp. 133-194. 44. Lubliner, J.; Oliver, J.; Oller, S.; and Onate, E., “A Plastic-Damage
20. Lormanometee, S., “Bond Strength of Deformed Reinforcing Bar Model for Concrete,” International Journal of Solids and Structures, V. 25,
under Lateral Pressure,” University of Texas at Austin, Austin, TX, Jan. No. 3, 1989, pp. 299-326. doi: 10.1016/0020-7683(89)90050-4
1974, 64 pp. 45. Hofstetter, G., and Meschke, G., Numerical Modeling of Concrete
21. Malvar, L. J., “Bond of Reinforcement under Controlled Confine- Cracking, Springer Science & Business Media, Berlin, Germany, 2011,
ment,” ACI Materials Journal, V. 89, No. 6, Nov.-Dec. 1992, pp. 593-601. 327 pp.

112 ACI Structural Journal/January 2018


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 115-S09

Seismic Performance of Innovative Reinforced Concrete


Coupling Beam—Double-Beam Coupling Beam
by Youngjae Choi, Poorya Hajyalikhani, and Shih-Ho Chao

Diagonally reinforced coupling beams (DCBs) are commonly used Two types of coupling beams are allowed by ACI 318-14
as seismic force-resisting members for medium- to high-rise build- (ACI Committee 318 2014) for coupled wall system. They
ings. The diagonal reinforcing bars in DCBs are most effective are conventional and diagonally reinforced coupling beams,
when the beam has a span-depth ratio of less than 2. However, and their application depends on the span-depth ratio (or
modern construction typically requires span-depth ratios between
aspect ratio ln/h) and shear demands. Prior studies (Barney et
2.4 and 4, which leads to a very shallow angle of inclination for
al. 1980; Naish at el. 2013; Lim et al. 2016) have shown that
the diagonal reinforcement. The lower angles of inclination, when
combined with the detailing requirements specified in ACI 318, conventionally reinforced coupling beams (reinforced with
result in reinforcement congestion and construction difficulties. vertical and longitudinal reinforcing bars) with span-depth
These issues can be considerably minimized by using an innova- ratios between 2.5 and 4 and maximum shear stress between
tive and simplistic reinforcing scheme consisting of two separate 3.4√fc′ and 7.9√fc′ exhibited fast strength degradation after
cages similar to those used for typical beams in reinforced concrete approximately 4% beam chord rotation. The major source of
special moment frames. The proposed coupling beam acts like a the strength degradation was the sliding shear at the beam-
conventional coupling beam under small displacements. Upon the to-wall interface.
occurrence of large displacements, cracks begin developing at the Another important factor when considering strength
beam’s midspan and midheight area where the narrow unreinforced degradation is rotational capacity. Prior nonlinear time-
concrete strip is located, gradually propagating toward the beam’s
history analyses (Harries and McNeice 2006) indicated that
ends. The cracks eventually separate the coupling beam into two
coupling beams would need average rotational capacities of
relatively slender beams where each has nearly twice the aspect
ratio of the original coupling beam. This split essentially trans- approximately 3% and 6% for design basis earthquakes
forms the shear-dominated deep beam behavior into a flexure-dom- (DBEs) and maximum considered earthquakes (MCEs),
inated slender beam behavior. Because damage initiates from the respectively. These rotational capacities help maintain the
center of the beam and then spreads toward the ends, the beam’s integrity of a coupled wall system. Note that DBE has a 10%
ends maintain their integrity even under very large displacements, probability of exceedance in 50 years and MCE has a 2%
thereby eliminating the sliding shear failure at the beam-to-wall probability of exceedance in 50 years.
interface. Testing results on half-scale specimens with span-depth Based on the shear resistance and adverse failure mecha-
ratios of 2.4 and 3.3 showed that the proposed coupling beam not nisms of conventional coupling beams, Paulay and Binney
only has high ductility and shear strength, but can significantly (1974) proposed diagonally reinforced coupling beams
reduce construction issues in conventional DCBs. In addition, the
(DCBs) with reinforcement detailing consisting of two
proposed coupling beam arrangement has great architectural flexi-
intersecting diagonal reinforcement groups combined with
bility, allowing utility ducts to be placed inside the coupling beams
where the gap between the two steel cages is located. closely spaced transverse reinforcement (Fig. 1(a)). In this
reinforcement detail, the diagonal bars need to be well
INTRODUCTION confined by transverse reinforcement and carefully anchored
Reinforced concrete (RC) structural walls are used in the walls. In a design using this type of coupling beam, the
commonly as the primary seismic-force-resisting system entire shear-transfer mechanism is resisted by heavily rein-
in buildings. Based on the architectural requirements, these forced diagonal cages. Experimental results have shown that
walls often have numerous openings for features such as diagonal reinforcement detailing can significantly improve
elevators, windows, and doors, which divide a single wall into deformation and energy dissipation capacity compared
more slender walls connected by substantial beams. These with conventional detailing for coupling beams subjected
beams are known as coupling beams. The use of the coupled to reverse cyclic loading (Barney et al. 1980; Galano and
wall system leads to a more efficient and economical structure Vignoli 2000; Naish et al. 2009; Paulay and Binney 1974).
system than single walls because properly designed coupled For coupling beams with a span-depth ratio of approxi-
wall systems possess significantly higher strength, stiffness, mately 1.0, diagonal reinforcement placed with a large incli-
and energy dissipation capacity. For the desired behavior of nation has proven to be the most efficient solution (Harries
the coupled wall system to be attained, the coupling beams are
required to sustain high shear forces while undergoing large ACI Structural Journal, V. 115, No. 1, January 2018.
MS No. S-2016-368.R3, doi: 10.14359/51700951, received March 10, 2017, and
displacement. However, the coupling beams must also yield reviewed under Institute publication policies. Copyright © 2018, American Concrete
before the wall piers, behave in a ductile manner, and exhibit Institute. All rights reserved, including the making of copies unless permission is
obtained from the copyright proprietors. Pertinent discussion including author’s
significant energy-dissipating characteristics. closure, if any, will be published ten months from this journal’s date if the discussion
is received within four months of the paper’s print publication.

ACI Structural Journal/January 2018 113


2005; Moehle et al. 2011; Naish et al. 2009; Moehle 2015).
Five prominent issues are listed as follows:
1. A small angle of inclination significantly decreases
the efficiency of diagonal reinforcement in resisting shear
forces, which in turn requires an even smaller angle to
accommodate the reinforcing bars. Thus, more reinforcing
bars are needed, which ultimately increases the difficulty
of construction. There is significant difficulty in placing
the diagonal reinforcing bars because they can be easily
obstructed by transverse reinforcement used to confine the
diagonal bars (Fig. 1(a)). Additional reinforcement detailing
is also required when the extension of the diagonal bars is
bent at the top of the wall or at openings in the wall.
2. Although ACI 318-14 Section 18.10.7.4(d) allows alter-
native full beam section confinement to reduce congestion
issues, there is still the obvious difficulty of passing the diag-
onal bars through the hoops and crossties (Fig. 1(b)).
3. It can be very challenging and time-consuming to thread
the diagonal reinforcement through the congested vertical
and horizontal bars in the wall’s boundary elements
(Fig. 1(a) and 1(b)).
4. The minimum width requirement for diagonal elements
causes interlock of the two diagonal elements. This in turn
demands an increased clear distance between reinforcing
bars in order for one diagonal element to pass through the
other. Often, the diagonal reinforcing bars need additional
space to prevent the conflict with the boundary element rein-
forcement of the wall pier. The minimum dimensions and
required reinforcement clearances can make the coupling
beam very wide, which controls the wall width.
5. Modern construction also calls for the additional archi-
tectural requirement of having utility ducts go through the
coupling beams (Hooper 2014). This is generally a diffi-
cult task when using diagonally reinforced coupling beams
(DCBs) due to reinforcement congestion or the concern of
Fig. 1–Reinforcement detail: (a) DCB with aspect ratio weakening the local region when ducts are present.
greater than 2.0, according to ACI 318-14 (confinement of Other alternative reinforcement schemes have been inves-
individual diagonals); (b) DCB with aspect ratio greater tigated (Tassios et al. 1996), such as the addition of dowels
than 2.0, according to ACI 318-14 (full confinement of diag- at the ends of the coupling beams or a diagonal reinforce-
onally reinforced concrete beam section); and (c) proposed ment located only at the beam-wall interface. However, it
reinforcement detailing (DBCB) for RC coupling beam. was experimentally demonstrated that coupling beams with
these alternative reinforcement details did not exhibit satis-
et al. 2005). On the other hand, modern architectural
factory seismic behavior, and/or they posed construction
specifications typically require span-depth ratios between
difficulties (Tassios et al. 1996).
2.4 and 4. Prior research shows that for DCBs with span-
The aforementioned construction and design issues with
depth ratios between 2.4 and 3.3, the majority of damage
DCBs can be considerably minimized by using an innova-
occurred at the beam-to-wall interface due to slip of the
tive and simplistic reinforcing scheme investigated in this
diagonal reinforcing bars from adjacent walls (Naish et al.
study, as illustrated in Fig. 1(c). This new coupling beam—
2013). This type of bond failure is typically difficult to repair
double-beam coupling beam (DBCB)—consists of two
(Engindeniz et al. 2005). In addition to the slip, damage at
separate cages similar to those used for beams in reinforced
the DCB ends revealed full-depth cracking, which leads to
concrete (RC) special moment frames. The new coupling
separation of the beam and wall (Paulay 1977). In addition,
beam acts like a conventional coupling beam under small
coupling beams with span-depth ratios between 2.4 and 4
displacements. With larger displacements, cracks begin
lead to a very shallow angle of inclination for the diagonal
developing at the beam’s midspan and midheight where
reinforcement (can be as low as approximately 10 degrees).
the narrow unreinforced concrete strip is located, gradually
The lower angles of inclination, combined with the detailing
propagating toward the ends. The cracks eventually separate
requirements specified in ACI 318-14, can cause several
the beam into two relatively slender beams (ACI 318-14,
major issues for both design and construction (Harries et al.
Section 18.6.2.1, defines a beam with clear span ln extending
at least four times the effective depth d as a relatively slender

114 ACI Structural Journal/January 2018


member) where each has nearly twice the aspect ratio of
the original coupling beam. This essentially transforms the
shear-dominated behavior into flexure-dominated behavior
common to conventional beams. Because damage initiates
from the center of the beam, then spreads toward the ends,
the beam ends maintain their integrity even under very
large displacements, thereby eliminating the sliding shear
failure at the beam-to-wall interface, as is commonly seen
in conventional beams (Aktan and Bertero 1981; Aristiza-
bal-Ocfaoa 1987). In addition, because the cracks always
initiate at midspan and midheight, the damage location can
be easily predicted, which makes repair work, if needed,
much easier after moderate earthquakes. Furthermore, the
gap between the two cages allows for utility ducts through
the beam without affecting the beam’s seismic behavior.
For performance-based design, a structural component’s
damage at various performance levels needs to be investi-
gated to evaluate its potential impact on life and economic
loss, down time, and occupancy. Post-earthquake damage
Fig. 2—Force-resisting mechanism in DBCBs.
states at various performance levels for DCBs, as outlined by
ACI 374.2R (ACI Committee 374 2013) and ASCE/SEI 41 reinforced coupling beams with span-depth ratios between
(ASCE 2014), were used to evaluate the performance of 2.4 and 3.3.
DBCBs. The three levels of performance are:
• Immediate Occupancy Performance Level (defined EXPERIMENTAL PROGRAM
as the post-earthquake damage state in which a structure Force-resisting mechanism of DBCBs
remains safe to occupy and essentially retains its When a DBCB is subjected to small shear forces, the
pre-earthquake strength and stiffness): Coupling beams vertical shear stress distribution along the beam height
experience diagonal cracking, and the maximum accept- follows that predicted by elastic theory. Associated hori-
able beam rotation is 0.6% rad. zontal shear stress also acts along the longitudinal axis of the
• Life Safety Performance Level (defined as the beam. Both stresses are the highest at the midheight of the
post-earthquake damage state in which a structure has beam. Thus, maximum shear strain takes place at midheight.
damaged components but retains a margin against the Because no shear reinforcement is present along the beam’s
onset of partial or total collapse): Coupling beams have midheight, concrete at this location cracks first. The shear
extensive shear and flexural cracks with some crushing, deformation gradually breaks the concrete and a gap
but concrete generally remains in place. The maximum develops from the midspan toward the ends, and eventually
acceptable beam rotation is 1.8% rad. separates the DBCB into two relatively slender beams. Note
• Collapse Prevention Performance Level (defined as that although the shear stress is theoretically the same at
the post-earthquake damage state in which the building beam ends (Section A in Fig. 2), the restraint imposed by the
is on the verge of partial or total collapse. Substantial adjacent walls prevents the DBCB from separation at the
damage to the structure has occurred, potentially ends. Therefore, the separation starts at midspan (Section B
including significant degradation in the stiffness and in Fig. 2), then propagates toward both ends as displacement
strength of the lateral-force-resisting system): Coupling increases. This damage process is opposite from that of
beams are shattered and virtually disintegrated. The conventionally reinforced coupling beams in which the
maximum acceptable beam rotation is 3% rad. damage starts from the beam-to-wall interface, thereby
drastically reducing the ability of transferring shear to the
RESEARCH SIGNIFICANCE walls and eventually leading to sliding shear failure or bond
This research investigated the seismic performance of RC slip failure.
double-beam coupling beams (DBCBs) used in RC coupled
core walls (CCWs). Innovative configurations were used Reinforcement details
that can significantly improve and simplify constructibility, Reinforcement details of the test DBCBs are illustrated in
design, and repair work. The new coupling beams also Fig. 3. Each DBCB has two separate steel cages, and each
exhibited excellent strength, stiffness, and ductility under consists of three layers of vertically distributed longitudinal
large displacement reversals. The reinforcement layout of reinforcing bars. Each steel cage is confined by closed trans-
the new DBCBs is similar to that used for typical beams in verse reinforcement. The advantages of using distributed
reinforced concrete special moment frames. Results from longitudinal reinforcing bars are: 1) they delay the propa-
this research show how the new detailing approach provides gation of flexure-shear cracks from becoming the domi-
simplified construction. DBCBs also have seismic behavior nating cracks before the gap opens and separates the DBCB.
equivalent or superior to that provided by the diagonally The diagonal cracks resulting from cyclic shear forces can
rapidly diminish the concrete’s shear resistance (Park and

ACI Structural Journal/January 2018 115


Fig. 3—Reinforcement details of specimens. (Note: Dimensions are in inches; 1 in. = 25.4 mm.)
Table 1—Specimen information
Transverse reinforcement†
Unreinforced ρplastic hinging, % ρnon-plastic hinging, % ρl,‡ % (main bar,
Specimen ln/h ln/d concrete strip, in. Vnormalized* (bar size, spacing) (bar size, spacing) intermediate bar) fcm,§ ksi
R2.4-SC-1 2.4 5.76 1 8.34 4.9 (No. 4, 1.5 in.) 2.9 (No. 4, 2.5 in.) 2.4 (No. 6, No. 6) 5.7
R2.4-NC-1 2.4 5.76 1 8.34 4.9 (No. 4, 1.5 in.) 2.9 (No. 4, 2.5 in.) 2.4 (No. 6, No. 6) 6.3
R3.3-SC-1 3.3 7.98 1 7.47 4.9 (No. 4, 1.5 in.) 1.8 (No. 4, 4.0 in.) 3.2 (No. 7, No. 7) 5.7
R2.4-SC-0.25 2.4 5.49 0.25 10.1 4.9 (No. 4, 1.5 in.) 2.9 (No. 4, 2.5 in.) 3.0 (No. 7, No. 6) 6.0
R2.4-SC-2-P 2.4 6.33 2 9.43 3.7 (No. 3, 1.0 in.) 2.6 (No. 3, 1.4 in.) 3.5 (No. 7, No. 6) 4.0
*
Normalized shear stress: Vnormalized = Vdesign /( f c′bw h) , where fc′ = 5 ksi (design compressive strength).

Transverse reinforcement ratio ρ is calculated by Av/(sbw). Here, Av is total transverse reinforcement area within transverse reinforcement spacing s; and bw is width of beam. Note
that hoops for R2.4-C-1, R2.4-NC-1, and R2.4-SC-0.25 are overlapped by two C-shaped stirrups so that area of one leg is two times original area of No. 3 stirrup.

ρl is reinforcement ratio of either top or bottom longitudinal bars of each individual beam; ρl = As/(bwd). Intermediate bars are not included in ratio.
§
fcm is measured concrete compressive strength.
Notes: 1 in. = 25.4 mm; 1 ksi = 6.895 MPa.

Paulay 1975); 2) vertically distributed longitudinal rein- 3.3. The ratios were selected to reflect modern architectural
forcing bars form a pattern of multiple finer cracks, which specifications, which typically require coupling beams with
helps in the development of a more uniform distribution of span-depth ratios of approximately 2.4 for residential build-
stresses along the transverse reinforcement (Paulay 1969); ings and 3.3 for office buildings. Ratios were also selected to
3) vertically distributed longitudinal reinforcing bars can compare the test results of DBCB specimens with previous
prevent sliding shear failure near the beam-to-wall bound- research on DCBs with the same span-depth ratios (Naish
aries (Xiao et al. 1999); and 4) a beam section with vertically et al. 2009). However, instead of using a beam width of
distributed longitudinal reinforcing bars can add additional 12 in. (305 mm) as is the case with DCB specimens (Naish
flexural strength and ductility (Wong et al. 1990). Therefore, et al. 2009), the width of DBCBs was reduced by 50% (6 in.
the width of the coupling beam can be reduced and wider [152 mm]) because this width is sufficient to accommodate
spacing between reinforcing bars can be used in the top of longitudinal reinforcement. Also, this smaller width can
the beam during construction for placing and controlling impose high shear stress on the DBCBs. In addition, the
vibration of concrete. development length of the longitudinal reinforcement is
only about 50% (for No. 7 bars) to 60% (for No. 6 bars)
Specimen design of that required by ACI 318-14 Section 18.8.5.3(b). This is
Five half-scale DBCBs were fabricated and tested under because the beam-to-wall boundary did not suffer severe
reversed cyclic loads. The information of each DBCB is damage (to be discussed later). This is opposite to DCBs,
shown in Table 1. The name of each specimen was desig- where the major damage is at the beam-to-wall boundary
nated according to test variables. Primary test variables were due to the slip and extension of the diagonal reinforcing bars
span-depth ratios, displacement (loading) protocol, and the (Naish et al. 2013).
size of the unreinforced concrete strip. For instance, Spec- All specimens have a 1 in. (25.4 mm) unreinforced
imen R2.4-SC-1 represents a DBCB with a span-depth ratio concrete strip (clear distance between the ends of the trans-
of 2.4 (R2.4) and a 1 in. (25.4 mm) unreinforced concrete verse reinforcement) at midheight, except for Specimens
strip (1), tested under symmetrical cyclic (SC) displacement R2.4-SC-0.25 and R2.4-SC-2-P (with PVC pipes going
protocol. Specimen R3.3-SC-1 has a span-depth ratio of through the gap), which had a 0.25 in. (6.35 mm) and 2 in.

116 ACI Structural Journal/January 2018


(51 mm) unreinforced concrete strip, respectively. The sizes a smaller probable beam end moment and shear demand.
of the unreinforced concrete strip were selected based on It was observed in this study’s preliminary testing that the
preliminary sizes taken from a nonlinear finite element maximum compression reinforcement strain reached 0.008
analysis conducted by VecTor2 (Wong and Vecchio 2002), without concrete crushing. This is because a large amount
which is explained later in this paper. The analysis shows that of confinement was provided by closely placed transverse
a wide unreinforced concrete strip could lead to a reduced and longitudinal reinforcement. Therefore, a maximum
moment arm for each cage; consequently, its moment capacity usable concrete strain of 0.006 was used for the calculation
as well as the overall shear strength of the DBCB is reduced. of the moment capacity of DBCBs, which leads to a higher
On the other hand, if the unreinforced concrete strip is too maximum probable moment. It should be also noted that the
narrow, the beam might not separate into two beams before the restraint from the adjacent walls can provide non-negligible
major shear cracks dominate the behavior, as in conventionally compression, which in turn increases the bending strength of
reinforced coupling beams (using only vertical and longitudinal the beam. Then, the expected peak shear stress (design shear
reinforcement) and can lead to early strength degradation. stress Vdesign) can be calculated from the sum of maximum
When a DBCB separates into two relatively slender probable beam end moments divided by the beam’s span.
beams, the individual beams will have a span-depth ratio The normalized peak shear stress of each specimen is calcu-
that meets ACI Code criteria for beams in special moment lated using the expected peak shear stress divided by the
frames—that is, the span of the beam shall be at least four square root of design concrete compressive strength (5 ksi
times greater than that of the effective depth of the beam [34.4 MPa]). These shear stress calculation results are listed
(ACI 318-14 Section 18.6.2.1). Therefore, each half of the in Table 1. However, the required transverse reinforcement
DBCB specimens was designed according to ACI Code of each specimen was controlled by ACI 318’s confinement
except when enforcing the requirement for longitudinal requirements rather than by shear demand. Except for Spec-
reinforcement ratio, which is limited up to 2.5% for beams imen R2.4-SC-2-P, all specimens used No. 3 transverse
of special moment frame. The top or bottom reinforcement reinforcement at a spacing of 1.5 in. (38 mm) in the plastic
ratio of each separated beam was slightly higher than 2.5% hinging region, which was assumed to have a length equal
(Table 1). This was done to increase the shear stress of the to the height of the DBCB. The first two hoops from the
specimens. As shown in Fig. 3, reinforcement details of beam-to-wall interfaces were placed with a spacing of 1.0 in.
Specimens R2.4-SC-1 and R2.4-NC-1 were identical, and (25.4 mm). This was done to increase confinement to the
the only difference between these two specimens was the concrete where the maximum moments occurred. Beyond
applied loading protocols. For these two specimens, No. 6 the plastic hinging region, the spacing of transverse rein-
bars (284 mm2) were used for both the main and interme- forcement was 2.5 in. (63.5 mm) for all specimens except
diate longitudinal reinforcement. In Specimen R3.3-SC-1, for Specimen R3.3-SC-1, which had 4 in. (101.6 mm) of
because this specimen had greater span length, No. 7 bars spacing. A wider spacing in Specimen R3.3-SC-1 was used
(387 mm2) were used for longitudinal reinforcement to considering the smaller shear demand due to the slender
increase the shear stress to equal that in Specimens R2.4- aspect ratio. Except for Specimen R2.4-SC-2-P, the closed
SC-1 and R2.4-NC-1. A pilot analytical study using VecTor2 hoops were made of two overlapping U-shaped stirrups by
indicated that if the unreinforced concrete strip is too narrow, welding (Fig. 3). Therefore, clear spacing of the transverse
the beam might not separate into two beams. To verify the reinforcement was less than using one single No. 3 hoop. For
analysis result, Specimen R2.4-SC-0.25 was fabricated with Specimen R2.4-SC-2-P, No. 3 hoops with a 135-degree hook
a 0.25 in. (6.35 mm) unreinforced concrete strip. This spec- were spaced more closely together (1 in. [25.4 mm]) than the
imen was also used to gauge the maximum shear stress and other specimens. The plastic hinging length was assumed as
rotation a DBCB can carry when it is not separated. No. 7 half of the beam height. Beyond the plastic hinging region,
bars (387 mm2) and No. 6 bars (284 mm2) were used as the 1.4 in. (35.6 mm) spacing was used. For all specimens, 2.5 in.
beam’s main and intermediate longitudinal reinforcement, (63.5 mm) spacing was used at the midspan to help the
respectively. Specimen R2.4-SC-2-P used the same bars as initiation of the separation.
used in R2.4-SC-0.25. However, the nominal shear strength
of Specimen R2.4-SC-2-P was lower than that of Specimen Specimen construction
R2.4-SC-0.25. This is because of the decreased moment arm Each specimen consisted of a coupling beam as well as
caused by using the 2 in. (50.8 mm) unreinforced concrete a pair of large and small reinforced concrete blocks repre-
strip. Specimen R2.4-SC-2-P aimed at reaching a shear stress senting adjacent structural walls. Laying on one side on the
demand of approximately 9.5√fc′ (psi), which is slightly strong floor, concrete was placed for all specimens. Form-
lower than the ACI limit on the maximum nominal shear works were made of plywood reinforced with 2 x 4 in. (51 x
strength of DCBs. The required transverse reinforcement 102 mm) lumber. DBCB formwork was inserted in between
was calculated based on the maximum probable moment at the formwork for the concrete blocks. Top and bottom cages
beam ends. An overstrength factor of 1.25 was multiplied to of the DBCB were built separately and were easily slid into
the specified yield strength of steel reinforcing bars when the cage of the concrete blocks (Fig. 4(a)). This construction
calculating the probable moment of all specimens. Note that process was much easier than that required for DCBs. Ready
following ACI 318’s assumption with a maximum useable mixed concrete was used to cast all specimens. The specified
concrete strain of 0.003 would lead to a small compres- concrete compressive and steel yield strength was 5000 and
sive stress in the compression steel, which in turn leads to 60,000 psi (34.5 and 414 MPa), respectively. The measured

ACI Structural Journal/January 2018 117


Fig. 4—(a) Construction photo; and (b) test setup.

Fig. 5—Loading protocols: (a) symmetrical loading protocol; and (b) near-collapse loading protocol.
average concrete compressive strengths fcm obtained on the specimens were subjected to cyclic loading in a displace-
testing dates are listed in Table 1. ment control mode, which produced predefined reversed
cyclic displacement patterns. Two loading protocols were
Test setup and instrumentation used, starting from a coupling beam chord rotation of
The specimens were cast horizontally, then rotated and 0.25% and reaching a maximum rotation of 12%. The first
placed in the test setup, with the large block fixed to the loading protocol consisted of symmetric cyclic (SC) loading
strong floor (Fig. 4(b)). Displacement reversals were applied using two to three cycles per deformation level (Fig. 5(a)).
via a vertical actuator, with the line of action of the actu- However, this type of loading is not representative of a
ator forces passing through the test specimen’s midspan to near-collapse (NC) earthquake response, which would be
produce an anti-symmetrical moment pattern in the coupling unsymmetrical and would contain fewer cycles of loading;
beam and zero moment at the beam’s midspan. The actu- hence, the loading protocol should contain displacements
ator was connected to the small block through a wide flange that are representative of the ratcheting effect that leads to
steel section. The load was transferred to the small block by structural collapse. Such a protocol was developed based
means of direct bearing and unbonded threaded bars passing on preliminary nonlinear analyses (Fig. 5(b)). Linear vari-
through the small block. Two steel links were used to provide able differential transformers (LVDTs) were placed at the
some moderate axial restraints for the beams because the boundary of the DBCB and the small block to measure beam
adjacent structural walls and surrounding slab would provide chord rotations and at the outer face of the large block to
non-negligible resistance to beam expansion upon cracking gauge its rotation and movement. Strains of transverse and
(Teshigawara et al. 1998; Lequesne 2011; Barbachyn et al. longitudinal reinforcement were measured by strain gauges
2012). Each steel link consisted of two channel sections and attached to the reinforcement.
was pin-connected to a T-stub. The T-stub was tightened to
each block by four high-strength threaded rods. To deter- EXPERIMENTAL RESULTS
mine the shear force carried by the links, one of the coupling Cracking and damage pattern
beam specimens was cut in the middle after testing and the Table 2 summarizes experimental test results for all spec-
loading block was displaced 4 in. (102 mm) upward. It was imens. The progressive cracking patterns and the failure
found that the maximum shear force carried by the links mechanisms are shown in Fig. 6. The specimens with a
was 44 lb (195 N), which is negligible compared with the 1 in. (25.4 mm) unreinforced concrete strip separated into
peak strength of the coupling beams. The coupling beam two beams that had nearly twice their original span-depth

118 ACI Structural Journal/January 2018


Table 2—Test results The cracking pattern of Specimen R2.4-SC-0.25 (with a
Specimen θyield,* % Vmax, kip Vmax/Vdesign θmax,% Vnormalized†
smaller unreinforced concrete strip height) was similar to the
other specimens in the early stage of loading. However,
R2.4-SC-1 2 68.5 1.29 8 10.1
the concrete crushing in the unreinforced concrete strip
R2.4-NC-1 2 89 1.68 11 12.5 progressed in a slower manner. Before the separation
R3.3-SC-1 2 68 1.43 8 10.0 occurred, a considerable number of flexural-shear cracks had
R2.4-SC-0.25 4 68 0.97 2 9.75
propagated across the entire depth of the beam, leading to a
drastically different shear stress distribution from that shown
R2.4-SC-2-P 2 63.5 1.06 6 11.16
in Fig. 2. In addition, horizontal cracks along the outermost
*
θyield is beam chord rotation when longitudinal bar yields. longitudinal reinforcing bars developed after considerable

Normalized shear stress Vnormalized is Vmax /( f cm bw h) , where fcm is measured compres- shear cracks occurred. This is believed to be due to dowel
sive strength. failure as many shear cracks developed, causing loss of the
Notes: 1 in. = 25.4 mm; 1 kip = 4.45 kN. aggregate interlock after the width of the cracks widened.
ratio. The cracking and damage patterns of the specimens Only the middle half of the unreinforced concrete strip sepa-
were very similar to one another. During the first positive rated, which made the portions above and below the strip to
cycle loading, diagonal tension cracks (shear cracks) were behave like beams with very short span-depth ratios. This
observed at the midspan and midheight of the specimens aggravated the shear-dominated damage in the separated
where the 1 in. (25.4 mm) unreinforced concrete strip region. Cover concrete came off in this region and the spec-
was located. As the load reversed, the same crack pattern imen eventually failed by shear dominated behavior
appeared at the same location, producing diagonal grids of (Fig. 6(a)). As shown in Fig. 6(e), Specimen R2.4-SC-2-P
cracks crossing each other. No flexural cracks were observed has a more complete separation toward the beam ends,
at that time. As the load increased, those cracks spread which was believed to be attributed to the presence of the
toward both ends of the specimens (Fig 6(a)). As cyclic utility pipes as well as the larger-size unreinforced concrete
loads continued increasing, concrete at the unreinforced strip. However, it had more shear cracks because Specimen
concrete strip was crushed due to the abrasion resulting R2.4-SC-2-P had higher shear stress demand and the transverse
from the sliding of the top and bottom beams. This crushing reinforcement ratio was less than the other specimens
initiated from midspan, progressing toward the beam ends, (Tables 1 and 2). These results indicate that the separation of
because the blocks provided the DBCBs with a confinement the unreinforced concrete strip is affected by the span-depth
effect near the boundaries. The relative displacement of ratio and the size of the strip. Additionally, placing the utility
the top and bottom beams due to the horizontal sliding is pipes at the ends of DBCBs seems to help the separation.
clearly seen by a marked grid on the specimen (Fig. 6(a)). At While a seismic-resistant coupling beam is expected to
1.5% chord rotation, diagonal cracks widened up to 0.16 in. experience severe damage at the Life Safety and Collapse
(4 mm). Canbolat et al. (2005) reported that diagonal cracks Prevention performance levels, it is expected to retain most
were first observed during the cycles at 0.25% rotation for of its pre-earthquake design strength with light damage
their DCB specimen, and the diagonal cracks widened up (ACI 374.2R; ASCE/SEI 41) at the Immediate Occupancy
to 0.12 in. (3 mm) at 1.5% rotation. This indicated that the Performance Level. For the Immediate Occupancy Perfor-
occurrence of initial cracks in DBCBs and their widths are mance Level, the maximum accepted beam chord rotation
similar to that of conventional DCBs. At 2% rotation, the end for DCBs is from 0.6% according to ACI 374.2R. Figure 6
of the separation is approximately 6 in. (152 mm) away from shows that at 0.75% rotation, DBCBs only had cosmetic
the beam-to-wall boundaries. Diagonal cracks at the end of concrete cracking and the peak strengths had not been
the separation spread toward the boundaries in a fan-shaped reached. Furthermore, the beam-to-wall interfaces were
pattern. Concrete crushing at all corners of the specimens undamaged. Therefore, DBCBs can be easily repaired by
was initiated at approximately 2% to 3% rotations. At this patching with new concrete to regain the elastic stiffness
time, the measured strain in longitudinal reinforcement after moderate earthquakes.
at the corners indicated that the reinforcement yielded in
compression. Beyond 4% rotation, the concrete crushing at Shear-versus-rotation responses
the corners caused all the cover concrete in the plastic region Shear force (and stress) versus beam chord rotation
to come off. Beyond 6% to 8% rotation, the core concrete responses for all specimens are illustrated in Fig. 7. For
crumbled due to the combined normal and shear stresses, comparison purposes, ACI 318’s upper-bound shear stress
leading to significant degradation in the shear resistance. Vu = ϕ10√fc′Acw is also plotted as dashed lines in Fig. 7. The
This was the primary reason for the strength drop of Speci- beam chord rotation was obtained by using the net vertical
mens R2.4-SC-1 and R3.3-SC-1. For Specimen R2.4-NC-1, displacements, which were measured between the two
which was subjected to the near-collapse protocol (Fig. 6(b)), beam-to-wall interfaces, excluding the fixed-end rotation or
the damage was not as severe as the damage observed in displacement, if any. This measurement was then divided
Specimen R2.4-SC-1. Note that all the specimens main- by the beam span. It is seen in Fig. 7(a) that Specimen
tained their integrity at the beam-to-wall interfaces up to R2.4-SC-1 was able to maintain a very high shear stress
the end of the tests, which effectively eliminated the sliding (10.1√fcm; fcm is the measured concrete compressive strength,
shear failure. or 10.8√fc′; fc′ is the design concrete compressive strength
= 5 ksi [34.5 MPa]) without significant strength degrada-

ACI Structural Journal/January 2018 119


Fig. 6—Cracking and damage pattern of DBCBs: (a) R2.4-SC-1; (b) R2.4-NC-1; (c) R3.3-SC-1; (d) R2.4-SC-0.25; and
(e) R2.4-SC-2-P. Note: These do not represent residual cracking (that is, at zero rotation).
tion up to a rotation of 6% (approximate demand for MCE previously. Specimen R2.4-SC-2-P retained a very high
level ground motions). Also, it could still resist 80% of the shear stress (11.16√fcm or 10√fc′) up to 6% rotation without
peak stress at 8% rotation. In addition, Specimen R2.4-NC-1 significant strength drop. However, after 6% rotation, its
showed no strength degradation up to 11% rotation, while shear strength dropped rapidly.
shear stress increased (12.5√fcm or 14√fc′) when subjected
to the near-collapse loading protocol, as shown in Fig. 7(b). Strains in steel reinforcement
Furthermore, the shear strength in Specimen R3.3-SC-1 did The yielding in the top and bottom layers of the longitu-
not drop until 8% rotation, maintaining a shear stress of dinal reinforcing bars of the two steel cages at the beam-to-
10√fcm (or 10.7√fc′) (Fig. 7(c)). The shear stress of Specimen wall interface indicates the complete separation of DBCBs
R2.4-SC-0.25 reached 9.75√fcm (or 10.7√fc′). However, it and a mechanical behavior shifting from shear to flexure. For
experienced significant shear strength degradation after 2% Specimens R2.4-RC-1, R2.4-NC-1, and R2.4-SC-2-P, the
rotation due to the shear dominated behavior, as discussed yielding of longitudinal reinforcing bars commenced at 2%

120 ACI Structural Journal/January 2018


Fig. 7—Hysteresis curves of DBCBs: (a) R2.4-SC-1; (b) R2.4-NC-1; (c) R3.3-SC-1; (d) R2.4-SC-0.25; and (e) R2.4-SC-2-P.
(Note: 1 in. = 25.4 mm; 1 kip = 4.45 kN; normalized measured shear force is Vtest /( f cm b w h) , where fcm is measured concrete
compressive strength.)

Fig. 8—Strain of longitudinal bar inside loading block (R2.4-SC-2-P). (Note: 1 in. = 25.4 mm.)
rotation, while it was 3% in Specimen R3.3-SC-1. For Spec- in the longitudinal reinforcement spread up to approximately
imen R2.4-SC-0.25, no yield in the longitudinal reinforcing 7.5 in. (191 mm) (which is 50% of the height of DBCBs)
bars was measured at the beam-to-wall interface. In addition, away from the beam-to-wall interface for Specimens R2.4-
strains in the top and bottom longitudinal reinforcing bars of SC-1, R-2.4-NC-1, and R2.4-SC-2-P, and approximately 15 in.
each individual cage at the beam-to-wall interface showed (38 mm) for Specimen R3.3-SC-1. This indicates that the
opposite signs, indicating no residual tensile strains affecting plastic hinging region of DBCBs depends on the span-depth
the behavior of DBCBs. This is because the reinforcing bars ratio. No yielding in transverse reinforcement was recorded
in DBCBs do not experience large inelastic deformation for Specimen R2.4-SC (NC)-1. However, only one hoop in
even at large displacement. On the other hand, due to the the plastic hinge region of Specimens R3.3-SC-1 and R2.4-
residual tensile strains after the reinforcing bars experience SC-2-P experienced yielding at 4% and 6% rotation, respec-
large inelastic deformation, both the top and bottom rein- tively. In addition, all intermediate longitudinal reinforce-
forcing bars in conventionally reinforced coupling beams ment experienced yielding, thereby contributing toward the
are usually subject to tension at large displacement. This strength of the DBCBs. Figure 8 illustrates the strain distri-
does not allow the interface crack to be entirely closed when bution from four strain gauges on a longitudinal reinforcing
subject to compression, thereby causing degradation of bar inside the concrete block (representing the wall pier) of
shear-transfer mechanism and sliding shear failure. Yielding Specimen R2.4-SC-2-P. As shown in the figure, strains in

ACI Structural Journal/January 2018 121


Fig. 9—Normalized shear stress versus beam chord rotation: (a) span-depth ratio of 2.4; and (b) span-depth ratio of 3.3. (Note:
Normalized shear stress is shear stress normalized by maximum shear stress of each specimen.)

Fig. 10—Comparison of DBCBs with DCBs: (a) span-depth ratio of 2.4; and (b) span-depth ratio of 3.3. (Note: Normalized
measured shear force is Vtest /( f cm b w h) , where fcm is measured concrete compressive strength.)
the reinforcing bars at the interface and at 4 in. (102 mm) In any case, a beam’s elastic stiffness can be easily adjusted
inside the block increased by a similar amount up to a 2% by optimizing its dimensions.
beam rotation. However, at larger rotations, only the strain
at the interface showed a dramatic increase. Although the Comparison between DBCB and DCB
specimen was subjected to many large displacement rever- The performance of DBCBs was compared with that of
sals, yielding did not penetrate beyond approximately 4 in. DCBs tested by Naish et al. (2009). Both DCB and DBCB
(102 mm) inside the wall pier. The embedded length of the specimens were approximately 1/2-scale replicas of the
longitudinal reinforcing bars in Specimen R2.4-SC-2-P is coupling beams in typical residential and office buildings.
18 in. (457 mm), which is only 50% of that required by The only major difference was that DBCBs were half the
ACI 318-14 Section 18.8.5.3(b), which is 37 in. (940 mm). width of DCBs. The design of the DCBs followed current
The minor yielding or complete elastic strain beyond 4 in. ACI provisions (ACI 318-14). Their DCBs can be divided
(102 mm) inside the wall pier indicates that DBCBs did into two groups: one group has transverse reinforcement
not suffer bond deterioration, as commonly seen in DCBs around diagonal bar groups (CB24D or CB33D), and the
(Naish et al. 2013). This is because, unlike DCBs, damage other group has transverse reinforcement around the entire
of a DBCB starts from the center of the beam and then cross section (CB24F or CB33F). Because the performance
gradually propagates toward the end. Figure 8 also shows of the latter group, in terms of strength and ductility, was
that, along the embedded length, the steel reinforcement slightly better (Naish et al. 2009), the performance of DBCBs
experienced completely elastic or minor inelastic cyclic is compared to CB24F and CB33F in Fig. 10(a) and (b). The
deformation up to approximately 3% beam rotation. shear stress, normalized by the square root of the measured
concrete compressive strength, fcm, was used for comparison
Stiffness because the specimens have different concrete compressive
Figures 9(a) and (b) illustrate the normalized shear stress strengths and cross-sectional areas. Two beam chord rota-
versus beam chord rotation response for DBCBs with the tions, 3% and 6%, are highlighted in the figures because they
symmetrical cycling load and DCBs tested in prior research represent approximately the upper bound rotational demands
(Naish et al. 2009). The normalized shear stress is the stress of coupling beams for the DBE and MCE level ground
divided by the maximum shear stress of each specimen. motions (Harries and McNeice 2006). The specimens with a
These figures indicate that DBCBs have an equivalent elastic span-depth ratio of 2.4 for both the DCBs and DBCBs
stiffness and act like a conventional coupling beam before exceeded a shear stress level of 10√fcm (psi), which is the
inelastic behavior occurs. In other words, DBCB’s elastic unfactored shear strength limited by the ACI provision.
stiffness is not affected by its unique separation mechanism. Although the strength in DBCB R2.4-SC-1 slightly

122 ACI Structural Journal/January 2018


Fig. 11—(a) Comparison of experiment and VecTor2 analysis; and (b) size effect of unreinforced concrete strip. (Note: 1 in. =
25.4 mm; 1 kip = 4.45 kN.)
decreased after 3% beam rotation, the overall ductility and Table 3—Material models
strength of DBCB R2.4-SC-1 were similar to that of the
Concrete models Concrete models
DCB specimen (CB24F) up to a beam rotation of 6% (MCE).
For the specimens with a span-depth ratio of 3.3, CB3.3F Compression Hognestad Crack width Agg/2.5 max.
prepeak (Parabola) check crack width
has a shear strength of 6.6√fcm (psi), and its shear strength
degradation begins after 3% beam rotation. On the other Compression Modified Crack slip Walraven
post-peak Park-Kent calculation (monotonic)
hand, DBCB R3.3-SC-1 could sustain much higher shear
stress (10√fcm [psi]) with no strength degradation up to 8% Compression
Vecchio 1992-A
Creep and
Not available
beam rotation. This shows that DBCBs can provide higher softening relaxation
safety by maintaining a stable global response of the struc- Tension Modified Bentz Hysteretic Nonlinear with
ture without loss of coupling between walls due to their high stiffening 2003 response plastic offsets)
plastic rotational capacity (Harries and McNeice 2006). Tension
Linear Concrete bond Eligehausen
stiffening
ANALYTICAL STUDY FRC tension Not considered — —
Nonlinear finite element analysis
Confined strength Kupfer/Richart Reinforcement
Nonlinear analysis using finite element software VecTor2
(Wong and Vecchio 2002) was conducted to study the effects Variable – Hysteretic Bauschinger
Dilation
Kupfer response Effect (Seckin)
of the width of the unreinforced concrete strip. A finite
element model was built for Specimen R2.4-SC-1. The trian- Cracking criteria
Mohr – Coulomb
Dowel action
Tassios
gular elements in VecTor2 were used for modeling concrete. (stress) (crack slip)
To simulate steel reinforcement, truss elements were used Crack stress Basic (DSFM/
Buckling
Refined Dhakal-
rather than smeared reinforcement. All the reinforcing bars calculation MCFT) Maekawa
simulated by truss elements were placed at the same posi-
tions as that in the actual specimen. A perfect bond between of the specimen in Fig. 11(a) is overestimated by VecTor2,
reinforcing bars and concrete was assumed. A smoothed the peak strengths and the beam rotations associated with the
triangle mesh was used with a mesh size of 0.63 in. (16 mm) starting of strength degradation reasonably match the exper-
based on a convergence study. The loading block and the imental results. It should be noted that VecTors2 stopped
large block were assumed to be rigid elements. Because running when severe cracking occurred at the unreinforced
VecTor2 does not provide any type of hinge boundary, the concrete strip location, although the strength of the DBCB
steel links, which connect the loading block to the large model did not drop significantly.
block for restraining the rotation of the loading block and
beam elongation, were omitted. Instead, to prevent the block Effect of height of unreinforced concrete strip
from rotating, displacement was imposed evenly on the top As shown in the experimental results, whether the DBCB
of the blocks. The material models used in this study are separates or remains unseparated depends on the height
shown in Table 3. A detailed explanation of the models’ of the unreinforced concrete strip. To assess the impact of
properties and functions can be found in the VecTor2’s User this factor, five models were studied. All models were the
Manual (Wong and Vecchio 2002). same in dimension and reinforcement detailing as Specimen
Figure 11(a) compares the simulation result from VecTor2 R24-SC-1, except they each had a different height unrein-
and the test result. The damage pattern from the analysis forced concrete strip. The model with a 0.25 in. (6.35 mm)
is very similar to the result of the experiment (refer to unreinforced concrete strip failed earlier than the others,
Fig. 11(a)), showing the beam separation and plastic hinges exhibiting the highest strength but poor ductility. On the
at the two separated beams (Fig. 11(a)). Comparison for other hand, the model with a 2 in. (51 mm) unreinforced
hysteresis response up to approximately 4% beam chord concrete strip exhibited the highest ductility, but lowest
rotation is shown in Fig. 11(a). Even though unloading stiffness strength (due to the reduced moment arm of each beam;

ACI Structural Journal/January 2018 123


however, strength can be easily raised by using larger 1. The test results on half-scale coupling beam specimens
reinforcing bars with wider DBCBs). It was observed in the with a span-depth ratio of 2.4 and 3.3 showed that coupling
analytical study that elastic stiffness of the beam is nearly beams with the proposed reinforcement scheme were able
the same when the size of the unreinforced concrete strip to simultaneously sustain high shear stresses (10 to 12√fcm
changes. In addition, the strength of the models increased [psi]) and large rotations (6 to 11%) before strength degra-
as the height of the unreinforced concrete strip decreased dation occurred.
(Fig. 11(b)). An additional model with a 2 in. (51 mm) 2. Experimental results showed that both DBCB and ACI
unreinforced concrete strip was analyzed. This model used compliant diagonally reinforced coupling beams (DCBs)
No. 7 reinforcing bars to replace No. 6 longitudinal bars at with a 2.4 span-depth ratio were able to sustain a high shear
the top and bottom of each cage, thereby compensating for stress of 10√fcm (psi) up to 6% beam rotation (approximate
the reduced strength with larger strip height. The analysis rotational demand for MCE ground motion and Collapse
(Fig. 11(b)) shows that, this design was able to effectively Prevention Performance Level). However, for the speci-
increase the strength while maintaining a high ductility. mens with 3.3 span-depth ratio, DBCBs not only showed
Based on the experimental and analytical results, for each better ductility (up to 8% beam rotation) than the DCBs (up
individual beam in a DBCB, the span-depth ratio should be to 6% beam rotation), but also reached greater shear strength
at least 5.0 to ensure that the unreinforced concrete strip is (10√fcm versus 6.6√fcm [psi]). These improvements offered
large enough to allow the DBCB to separate. The depth of higher safety by maintaining the structure’s stable global
each beam was calculated as 0.5 × (depth of entire DBCB response without loss of coupling between walls.
minus height of the unreinforced concrete strip). 3. The beam-to-wall interface in DBCBs experienced
much less damage when compared to that of DCBs; there-
SUMMARY AND CONCLUSIONS fore, a smaller development length is required for the longi-
This study investigated an innovative and simplistic rein- tudinal reinforcing bars (approximately 60% of that required
forcing layout for RC coupling beams that can significantly by ACI 318-14 Section 18.8.5.3(b)).
reduce design and construction difficulties when using 4. Experimental and nonlinear FEA shows that the height
diagonally reinforced coupling beams. The new double- of the unreinforced concrete strip does not have a signifi-
beam coupling beam (DBCB) consists of two separate cant effect on the elastic stiffness of the DBCBs. However,
cages much like those used for typical beams in reinforced a smaller height (0.25 in. [6.35 mm]) could not separate the
concrete special moment frames. The two cages are sepa- two beams before the major diagonal cracks developed,
rated by a small spacing from 1 to 2 in. (25.4 to 50.8 mm). thereby reducing the ductility of the DBCBs. On the other
Only vertical and horizontal reinforcing bars are needed. hand, a large height will decrease the moment arm of each
Upon large displacements, cracks begin developing at a beam, leading to a smaller capacity. Nevertheless, this can be
DBCB’s midspan and midheight, then gradually propagate easily compensated for by using slightly larger reinforcing
toward the beam’s ends. The cracks eventually separate the bars or wider DBCBs with slightly more reinforcing bars.
coupling beam into two relatively slender beams where each 5. Because the cracks at DBCBs always initiate at their
has nearly twice the aspect ratio of the original coupling midspan and midheight, the damage location can be easily
beam. This split essentially transforms the shear-dominated accessed, which makes repair work easier after moderate
single deep beam behavior into a flexure-dominated slender earthquakes. Test results showed that at 1% beam rotation
beam behavior. Because damage initiates from the center of (rotational demand for Immediate Occupancy Performance
the beam, and then spreads toward the ends, the beam ends Level [ACI 374.2R]) DBCBs only had cosmetic concrete
are able to maintain their integrity even under very large cracking and the peak strengths had not yet been reached. The
displacements, thereby eliminating the sliding shear failure beam-to-wall interfaces were essentially undamaged. This
at the beam-to-wall interface. Although DCBs have excel- makes repair work easier after moderate earthquakes. This is
lent seismic performance, they are very difficult to construct opposite to DCBs, where the major damage is at the beam-
with a beam span-depth ratio higher than 3. On the other wall boundary as a result of the slip and extension of the diag-
hand, the proposed DBCBs possess simultaneously high onal bars (Naish et al. 2013), which is difficult to repair.
shear strength and ductility with a much simpler design and 6. Based on the experimental and analytical results, for each
construction process. The width of adjacent walls does not individual beam in a DBCB, the span-depth ratio should be at
have to be increased due to the required width of coupling least 5.0 to ensure that the unreinforced concrete strip is large
beams because the width of DBCBs can be largely reduced enough to allow the DBCB to separate. The depth of each beam
compared to DCBs. In addition, the placement of longitu- was calculated as 0.5 × (depth of entire DBCB minus height of
dinal reinforcing bars in DBCBs can be easily adjusted to the unreinforced concrete strip). A detailed design and modeling
accommodate the locations of vertical longitudinal rein- of DBCBs will be presented in future publications.
forcing bars in the wall pier’s boundary elements. Further-
more, diagonal bars in DCBs need to be bent at the top floor AUTHOR BIOS
in order not to protrude outside of the concrete wall. This ACI member Youngjae Choi is a PhD Student and Research Assistant in
the Department of Civil Engineering at the University of Texas at Arlington,
could make the construction even harder and affect the Arlington, TX. His research interests include design and behavior of rein-
force-transferring mechanism. On the other hand, DBCBs forced concrete structures subjected to extreme loads.
can easily resolve these issues.
Poorya Hajyalikhani is a Structural Engineer at Toll Brothers, Grapevine,
The following conclusions are drawn from the study: TX. He received his PhD in structural engineering from the University of

124 ACI Structural Journal/January 2018


Texas at Arlington. His research interests include the seismic behavior and Hooper, J. D., 2014, “Couplings—Current Practice, Issues, and Oppor-
design of reinforced concrete coupling beams and squat walls with innova- tunities,” ID ORAL20D, Tenth U.S. National Conference in Earthquake
tive reinforcement configurations. Engineering, Earthquake Engineering Research Institute, Anchorage, AK,
July 21-25.
ACI member Shih-Ho Chao is a Professor at the University of Texas at Lequesne, R. D., 2011, “Behavior and Design of High-Performance
Arlington. He is a member of ACI Committee 239, Ultra-High-Performance Fiber-Reinforced Concrete Coupling Beams and Coupled-Wall Systems,”
Concrete, and 544, Fiber-Reinforced Concrete. He received the ACI Chester PhD dissertation, Department of Civil and Environmental Engineering,
Paul Siess Award for Excellence in Structural Research in 2011. His research University of Michigan, Ann Arbor, MI, 277 pp.
interests include high-performance fiber-reinforced cementitious materials Lim, E.; Hwang, S. J.; Cheng, C. H.; and Lin, P. Y., 2016, “Cyclic Tests
and seismic behavior of reinforced concrete structures. of Reinforced Concrete Coupling Beam with Intermediate Span-Depth
Ratio,” ACI Structural Journal, V. 113, No. 3, May-June, pp. 515-524. doi:
10.14359/51688473
REFERENCES Moehle, J., 2015, Seismic Design of Reinforced Concrete Buildings,
ACI Committee 318, 2014, “Building Code Requirements for Struc- McGraw-Hill Education, New York, 760 pp.
tural Concrete (ACI 318-14) and Commentary (ACI 318R-14),” American Moehle, J. P.; Ghodsi, T.; Hooper, J. D.; Fields, D. C.; and Gedhada, R.,
Concrete Institute, Farmington Hills, MI, 520 pp. 2011, “Seismic Design of Cast-In-Place Concrete Special Structural Walls
ACI Committee 374, 2013, “Guide for Testing Reinforced Concrete and Coupling Beams: A Guide for Practicing Engineers,” NEHRP Seismic
Structural Elements under Slowly Applied Simulated Seismic Loads (ACI Design Technical Brief No. 6, NIST GCR 11-917-11REV-1, produced by
374.2R-13),” American Concrete Institute, Farmington Hills, MI, 18 pp. the NEHRP Consultants Joint Venture, NIST, Gaithersburg, MD, 37 pp.
Aktan, A. E., and Bertero, V. V., 1981, “The Seismic Resistant Design of Naish, D.; Fry, J. A.; Klemencic, R.; and Wallace, J., 2013, “Reinforced
R/C Coupled Structural Walls,” Report No. UCB/EERC-81/07, Earthquake Concrete Coupling Beams—Part I: Testing,” ACI Structural Journal,
Engineering Research Center, University of California, Berkeley, Berkeley, V. 110, No. 6, Nov.-Dec., pp. 1057-1066.
CA, 196 pp. Naish, D.; Wallace, J.; Fry, J. A.; and Klemencic, R., 2009, “Reinforced
Aristizabal-Ocfaoa, J. D., 1987, “Seismic Behavior of Slender Coupled Concrete Link Beams: Alternative Details for Improved Constructability,”
Wall Systems,” Journal of the Structural Division, ASCE, V. 113, No. 10, Oct., Report to Charles Pankow Foundation, UCLA-SGEL, 103 pp.
pp. 2221-2234. doi: doi10.1061/(ASCE)0733-9445(1987)113:10(2221) Park, R., and Paulay, T., 1975, Reinforced Concrete Structures, John
ASCE, 2014, “Seismic Evaluation and Retrofit of Existing Buildings Wiley & Sons, Inc., New York, 769 pp.
(ASCE/SEI 41-13),” American Society of Civil Engineers, Reston, VA, Paulay, T., 1969, “The Coupling of Shear Walls,” PhD dissertation,
518 pp. Department of Civil Engineering, University of Canterbury, Christchurch,
Barbachyn, S. M.; Kurama, Y. C.; and Novak, L. C., 2012, “Analytical New Zealand, 432 pp.
Evaluation of Diagonally Reinforced Concrete Coupling Beams under Paulay, T., 1977, “Ductility of Reinforced Concrete Shearwalls for
Lateral Loads,” ACI Structural Journal, V. 109, No. 4, July-Aug., Seismic Areas,” Reinforced Concrete Structures in Seismic Zones, SP-53,
pp. 497-507. N. M. Hawkins and D. Mitchell, eds., American Concrete Institute,
Barney, G. B.; Shiu, K. N.; Rabbat, B. G.; Fiorato, A. E.; Russell, H. Farmington Hills, MI, pp. 127-148.
G.; and Corley, W. G., 1980, “Behavior of Coupling Beams under Load Paulay, T., and Binney, J. R., 1974, “Diagonally Reinforced Coupling
Reversals (RD068.01B),” Portland Cement Association, Skokie, IL, 22 pp. Beams of Shear Walls,” Shear in Reinforced Concrete, SP-42, American
Canbolat, B. A.; Parra-Montesinos, G. J.; and Wight, J. K., 2005, “Exper- Concrete Institute, Farmington Hills, MI, pp. 579-598.
imental Study on Seismic Behavior of High-Performance Fiber-Reinforced Tassios, T. P.; Moretti, M.; and Bezas, A., 1996, “On the Behavior and
Cement Composite Coupling Beams,” ACI Structural Journal, V. 102, Ductility of Reinforced Concrete Coupling Beams of Shear Walls,” ACI
No. 1, Jan.-Feb., pp. 159-166. Structural Journal, V. 93, No. 6, Nov.-Dec., pp. 711-720.
Engindeniz, M.; Kahn, L. F.; and Zureick, A.-H., 2005, “Repair and Teshigawara, M.; Kato, M.; Sugaya, K.; and Matsushima, Y., 1998,
Strengthening of Reinforced Concrete Beam-Column Joints: State of the “Energy Absorption Mechanism and the Fluctuation of Shear Force in the
Art,” ACI Structural Journal, V. 102, No. 2, Mar.-Apr., pp. 187-197. Coupled Shear Walls,” Structural Engineering World Wide 1998—Proceed-
Galano, L., and Vignoli, A., 2000, “Seismic Behavior of Short Coupling ings, Paper Number T-186-5, Elsevier Science Ltd., 8 pp.
Beams with Different Reinforcement Layouts,” ACI Structural Journal, Wong, P. K. C.; Priestley, M. J. N.; and Park, R., 1990, “Seismic Resis-
V. 97, No. 6, Nov.-Dec., pp. 876-885. tance of Frames with Vertically Distributed Longitudinal Reinforcement in
Harries, K. A.; Fortney, P. J.; Shahrooz, B. M.; and Brienen, P. J., 2005, Beams,” ACI Structural Journal, V. 87, No. 4, July-Aug., pp. 488-498.
“Practical Design of Diagonally Reinforced Concrete Coupling Beams— Wong, P. S. and Vecchio, F. J., 2002, “VecTor2 & Formworks User’s
Critical Review of ACI 318 Requirements,” ACI Structural Journal, V. 102, Manual,” Aug.
No. 6, Nov.-Dec., pp. 876-882. Xiao, Y.; Esmaeily-Ghasemabadi, A.; and Wu, H., 1999, “High-Strength
Harries, K. A., and McNeice, D. S., 2006, “Performance-Based Design Concrete Short Beams Subjected to Cyclic Shear,” ACI Structural Journal,
of High-Rise Coupling Wall Systems,” Structural Design of Tall and V. 96, No. 3, May-June, pp. 392-400.
Special Buildings, V. 15, No. 3, pp. 289-306. doi: 10.1002/tal.296

ACI Structural Journal/January 2018 125


in
Your
Classroom
Integrate into
your classroom!
To support future leaders, ACI has launched several initiatives to engage
students in the Institute’s activities and programs – select programs that
may be of interest to Educators are:
• Free student membership – encourage • Scholarships and fellowships – students
students to sign up who win awards are provided up to $10,000
• Special student discounts on ACI 318 and may be offered internships and paid
Building Code Requirements for Structural travel to attend ACI’s conventions
Concrete, ACI 530 Building Code Require- • ACI Award for University Student Activities –
ments and Specification for Masonry receive local and international recognition
Structure, & Formwork for Concrete manual. for your University’s participation in concrete
• Access to Concrete International – free to related activities
all ACI student members • Free access to ACI Manual of Concrete
• Access to ACI Structural Journal and ACI Practice – in conjunction with ACI’s chapters,
Materials Journal – free to all ACI student students are provided free access to the
members online ACI Manual of Concrete Practice
• Free sustainability resources – free copies • ACI online recorded web sessions and
of Sustainable Concrete Guides provided to continuing education programs – online
universities for use in the classroom learning tools ideal for use as quizzes or
• Student competitions – participate in ACI’s in-class study material
written and/or team-based competitions

https://www.concrete.org/educatorsandresearchers/aciinyourclassroom.aspx
ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 115-S10

A New Approach to Modeling Tension Stiffening in


Reinforced Concrete
by Angus Murray, Raymond Ian Gilbert, and Arnaud Castel

A novel approach to the modeling of tension stiffening is proposed


based on considerations of the highly non-uniform strain regions
that occur in the concrete surrounding embedded reinforcement
bars near the primary cracks. A finite element study is undertaken to
determine the effect of these disturbed regions on the overall defor-
mation of reinforced concrete tension members. Simple hand calcu-
lations are presented to determine an effective cross-sectional area
of concrete, which describes the overall contribution to axial rigidity
of the cracked concrete for the ideal case of perfect bond. A scalar
damage parameter is then introduced to account for the additional
reduction in the tension stiffening effect caused by deterioration of
bond at the reinforcement-concrete interface. Experimental tension
stiffening studies in the literature (including a recent study by the
authors) are used to calibrate the evolution of interface damage
according to factors such as applied loading and shrinkage.

Keywords: bond; disturbed regions; finite element analysis; reinforced


concrete; tension stiffening.

INTRODUCTION
Load sharing approach
Under service loads, the stiffness of a reinforced concrete
(RC) member is strongly affected by the sharing of tension
forces between the reinforcement and the concrete within the
cracked region of the member (that is, the tension stiffening
effect). Accurate models of tension stiffening are therefore
crucial for the successful design of serviceable RC structures.
Analytical models based on the principle of load sharing
are prominent in the literature because they are conceptu-
ally simple and allow a variety of serviceability effects to
be predicted, often with hand calculations alone. The typical
approach employed by this type of model is to consider the
variations of stress and strain in reinforcement and concrete,
which are governed by assumptions of the magnitude and
distribution of bond stresses τb at the interface between the
two materials (Fig. 1(a)). In this way, the tension stiffening
effect may be quantified as the difference between the strain
in the bare reinforcement, εs2, and the average reinforcement
strain within the cracked region of the member, εsm. The
tension stiffening effect is illustrated in Fig. 1(b) for an RC
tension member subjected to a service load Ns. Fig. 1—(a) Distribution of bond stress τb and reinforcement
Various approaches have been taken in the literature to strain εs for a cracked RC tension member; and (b) load-
characterize the interface behavior between reinforcement deformation responses of cracked reinforced concrete and
and concrete. For example, Marti et al.1 adopted a rigid- bare reinforcement.
plastic bond model leading to a linear variation of strain in
the reinforcement and concrete between adjacent cracks. ACI Structural Journal, V. 115, No. 1, January 2018.
MS No. S-2016-404.R4, doi: 10.14359/51700952, received March 28, 2017, and
Good agreement was achieved with experimental results, reviewed under Institute publication policies. Copyright © 2018, American Concrete
although the success of this model depends largely on the Institute. All rights reserved, including the making of copies unless permission is
obtained from the copyright proprietors. Pertinent discussion including author’s
specific choice of bond stress, for which little justification closure, if any, will be published ten months from this journal’s date if the discussion
is received within four months of the paper’s print publication.

ACI Structural Journal/January 2018 127


is provided. A similar model was presented by Gilbert and
Ranzi,2 in which the magnitude of the bond stress was modi-
fied to account for the deterioration of tension stiffening that
occurs with increasing load. The effects of other parameters on
the magnitude of the bond stress were also modeled, including
bar diameter and the time-dependent deformation of concrete
due to creep and shrinkage. Alternatively, models based on
linear bond stress distributions (and therefore quadratic vari-
ations of strain) have also been used successfully to describe
the in-service behavior of RC structures.3-6
Assumptions of both uniform and linear bond distributions
broadly corroborate experimental measurements of the strain
variation along embedded reinforcement bars.7,8 However, the Fig. 2—Concrete axial stress and strain variation close to
shape of the bond stress distribution is distinctly nonlinear, and remote from a primary crack.
and at any point between cracks depends on the local state
of damage at the reinforcement-concrete interface. With cross sections should exhibit a gradual transition from
increasing load, regions of high local bond stress (particularly highly non-uniform deformation to essentially plane defor-
near the primary cracks) exhibit a softening effect, as internal mation (Fig. 2). For RC tension members like those consid-
cone-shaped cracks and longitudinal splitting cracks begin to ered in this paper, this transition occurs over a distance
form around the embedded bars.9,10 The resulting bond stress that is roughly in the order of the width of cross section16
distribution is both nonlinear and load-dependent, and this is (≈2c + db), where c and db are the cover dimension and
reflected in some models in the literature.11,12 reinforcement bar diameter, respectively. After a stabilized
Another type of approach is to adopt a local bond stress- crack pattern has formed, however, significant portions of a
slip (τb-s) relationship.13 For simple τb-s models, explicit member surrounding the cracks (if not the entire member)
solutions may be found to characterize the tension stiff- are likely to exhibit this warping effect; the usual assumption
ening effect.14 However, most common τb-s models are that plane sections should remain plane does not hold true
piecewise-defined, and neat analytical solutions do not exist after the formation of primary cracks.
to quantify the tension stiffening effect for cases in which Some important aspects of the in-service behavior of RC
multiple branches of the τb-δ curve must be considered simul- structures can only be explained by considering this behavior.
taneously. To this end, numerical approaches are usually One example is the curved surfaces of primary cracks (Fig. 2),
favored,15 although these require significant computational which account for the difference in the measured crack width
effort, particularly for the analysis of entire structures. at the surface of a member compared to that at the level of
Notwithstanding, the tension stiffening effect is almost reinforcement.17-19 This non-uniform deformation engenders
always presented as though it were strictly a one-dimensional a dependence of crack width on the cover dimension, and
problem, with variations of stress and strain being consid- this fact is reflected in many crack models in the
ered in the direction of the embedded reinforcement only. literature.2,20,21
While this is a convenient simplification for the basis of an The formation of internal cone-shaped cracks near the
analytical model, it necessarily implies that the deformation primary cracks9 is also largely a result of non-uniform
of concrete at a specific cross section may be represented concrete axial stress variations transverse to the reinforce-
by a single value of stress and strain. However, even at the ment. High bond stresses in the area of the primary cracks
service level, it is difficult to justify such an assumption. induce regions of high tensile stress in the concrete at the
reinforcement-concrete interface and cause cone-shaped
Plane sections do not remain plane cracks to form at regular intervals around the embedded
In the plane of a primary crack, the entire tension force is reinforcement. Near the surface of the member, however, the
carried by the reinforcement and the concrete is free of stress concrete stress is small (Fig. 2) and may even be compres-
and strain. Through the action of bond, a portion of the total sive.17,19 In these regions near the primary cracks, prop-
tension force is gradually transferred from the reinforcement agation of the internal cracks is effectively arrested and
to the surrounding concrete as the distance from the crack they typically remain confined within the cover concrete.
increases. The difference between the bare reinforcement strain Conversely, regions remote from existing primary cracks are
and the average reinforcement strain between the primary characterized by essentially plane deformation and there-
cracks (εs2 – εsm) represents the tension stiffening effect. fore allow for the formation of new primary cracks, which
However, irrespective of the actual bond stress distri- (unlike internal cracks) occur almost instantaneously over
bution, the resultant bond force acting upon the concrete the entire cross section.
occurs over a region that is many times smaller than the Finally, for RC tension members tested in the laboratory
dimensions of the concrete cross section. This highly local- (whose ends represent initial primary cracks), there is a
ized loading induces a warping effect in the concrete cross significant and measurable difference in stiffness compared
sections, particularly near the primary cracks where bond to the ideal uncracked case, which may be observed during
stresses are relatively large. Over a sufficient distance from the very first application of load. During the earliest stages
a crack, Saint-Venant’s principle indicates that the concrete of loading, bond is mainly assured through a combination of

128 ACI Structural Journal/January 2018


chemical adhesion and micromechanical interlock between strain εsm that is of primary interest because it determines
the reinforcement and the concrete,10 and a condition of the overall tension stiffening effect. Therefore, an effective
“perfect bond” occurs everywhere (that is, no slip occurs along axial rigidity RA,ef is introduced that defines the relationship
the reinforcement-concrete interface). However, a condition between an axial force and the average reinforcement strain
of perfect bond is incompatible with a state of plane defor- (that is, εsm = N/RA,ef). In general, the relationship between
mation near a primary crack because it necessarily requires the effective and the local axial rigidity is given by
the concrete in this region to sustain a uniform tensile stress
over its entire cross section. The concrete is essentially free of 1 1 b 1
stress and strain close to the face of a crack. This incompat- = ∫ dz (3)
RA, ef b − a a RA ( z )
ibility may only be overcome if the concrete should exhibit
non-plane deformation near a primary crack.
where RA(z) is the local axial rigidity between some points
RESEARCH SIGNIFICANCE a and b along the reinforcement. Because RA(z) is bounded
The model proposed in this paper provides a more by the limiting values of RA2 and RA1, it follows from Eq. (3)
informed picture of the influence of non-plane deforma- that RA2 ≤ RA,ef ≤ RA1, as expected. The effective axial rigidity
tion occurring near primary cracks on the stiffness of RC RA,ef of a cracked RC tension member may be considered to
structures under service loads. The approach outlined in be governed by two primary effects, namely: 1) the spacing
this paper may be extended to characterize other aspects of primary cracks; and 2) the deterioration of bond at the
of serviceability in RC structures, including cracking and reinforcement-concrete interface. In this paper, the influ-
time-dependent effects. ences of these two effects on RA,ef are considered separately.
First, the individual effect of primary crack spacing on
CHARACTERIZATION OF DEFORMATION RA,ef may be highlighted by considering the special case of
Local axial rigidity perfect bond, for which no slip occurs at any point along
For an RC tension member subjected to an axial force the reinforcement-concrete interface. Perfect bond is
N, the reinforcement strain εs at a specific cross section is assured during the earliest stages of loading, during which
governed by the local axial rigidity RA (that is, εs = N/RA). At the primary bond mechanism is a combination of chemical
a point that is sufficiently remote from the abrupt disconti- adhesion and micromechanical interaction between the rein-
nuities in loading (bond stresses) and geometry presented by forcement and the concrete.10 This condition of zero slip is
a primary crack, a condition of essentially plane deforma- significant because it enforces the greatest possible interac-
tion occurs over the entire cross section and the local axial tion between the reinforcement and the concrete, and there-
rigidity is equal to RA1 = EsAs + EcAc, where Es and Ec and fore yields the greatest tension stiffening effect for a given
As and Ac are the elastic moduli and cross-sectional areas of spacing of primary cracks. For the case of perfect bond,
the reinforcement and concrete, respectively. This leads to the effective axial rigidity may be written as RA,ef = EsAs +
the familiar expression for the reinforcement strain in the EcAc,ef, where Ac,ef (0 ≤ Ac,ef ≤ Ac) is an effective cross-sec-
uncracked concrete, εs1 tional area of concrete.
Where the spacing of primary cracks is relatively wide,
most of a member’s cross sections are located in regions
N N
ε s1 = = (1) characterized by plane or nearly plane deformation with local
RA1 Es As + Ec Ac axial rigidity RA ≈ RA1. The cross sections near the primary
cracks, having local axial rigidity RA ≈ RA2, represent only a
In the plane of a primary crack, the concrete carries no tension small portion of the member’s total length. As a result, the
and the local axial rigidity is simply that of the bare reinforce- member’s effective axial rigidity RA,ef is close to (but less
ment (RA2 = EsAs), leading to a reinforcement strain εs2 of than) that of an uncracked member and the effective cross
sectional area of concrete, Ac,ef, is close to (but less than) the
N N actual cross-sectional area of concrete, Ac. Conversely, where
εs2 = = (2)
RA 2 Es As primary cracks are finely spaced, the concrete carries little of
the total tension force, and most cross sections exhibit highly
non-uniform deformation with local axial rigidity RA ≈ RA2.
In the intermediate region between a primary crack and a
In this case, RA,ef is close to (but greater than) that of the
cross section characterized by plane deformation, a gradual
bare reinforcement and the effective cross-sectional area of
transition must occur in the concrete’s contribution to local
concrete, Ac,ef, is close to (but greater than) 0.
axial rigidity (that is, from 0 to EcAc). At any point between
However, only a modest increase in load is required to
adjacent cracks, this contribution determines the extent to
break the chemical and micromechanical bonds that assure a
which the total axial force N is shared between the reinforce-
condition of perfect bond, and slip between the reinforce-
ment and the concrete, and, therefore, governs the variations
ment and the concrete facilitates a transition of the bond
of stress and strain in both materials.
mechanism to one characterized by bearing of the reinforce-
ment lugs on the cover concrete. With increasing load, internal
Effective axial rigidity
cone-shaped cracks and splitting cracks begin to form along
Whereas the local reinforcement strain εs is governed by
the interface, and this is accompanied by a reduction in
the local axial rigidity RA, it is the average reinforcement

ACI Structural Journal/January 2018 129


the effectiveness of the bond.10 Eventually, damage to the
reinforcement-concrete interface becomes so extensive that
the tension stiffening effect is exhausted, and RA,ef approaches
RA2, regardless of the spacing of primary cracks. The influence
of bond deterioration may be captured by a scalar damage
parameter λ (0 ≤ λ ≤ 1), which modifies the contribution to
axial rigidity of the effective area of concrete EcAc,ef. This
approach is similar to studies by Castel et al.,4 Xu et al.,5 and
Murray et al.6 The case of perfect bond is represented by λ = 0,
whereas complete bond deterioration is represented by λ = 1.
Thus, for the general case in which both primary crack spacing
and bond deterioration are considered, the effective axial
rigidity may be expressed as RA,ef = EsAs + (1 – λ)EcAc,ef and
the average reinforcement strain is given by

N N
ε sm = = (4)
RA, ef Es As + (1 − λ ) Ec Ac , ef

Together, Ac,ef and λ describe the two major mechanisms of


the tension stiffening effect in a cracked RC member. In the Fig. 3—(a) Segment of a cylindrical RC tension member; (b)
following sections, a finite element (FE) approach is used to schematic diagram of axisymmetric FE model.
determine Ac,ef for a wide variety of section geometries and
average primary crack spacings. The results of experimental 1
ρef = (6)
tension stiffening studies in the literature (including a recent n(ε s 2 /ε sm − 1)
study by the authors) are then used to calibrate the evolution of
λ according to factors such as applied loading and shrinkage. where ρef = As/Ac,ef is the effective reinforcement ratio; n =
Es/Ec is the modular ratio; and εs2/εsm is the ratio between
FINITE ELEMENT MODELING OF EMBEDDED the bare reinforcement strain εs2 and the average reinforce-
REINFORCEMENT ment strain of an embedded bar εsm, to be determined by
In the previous section, the effective cross-sectional area FE analysis.
of concrete, Ac,ef, was used to describe the influence of A schematic diagram of the FE model is shown in
primary cracks alone on the tension stiffening effect in RC Fig. 3(b). For the case of λ = 0, the dominant bond mecha-
tension members—that is, in the absence of any bond deteri- nism is a combination of chemical and physical adhesion,
oration. Accordingly, Ac,ef is determined for the special case although some portion of the total bond may be attributed to
of perfect bond (λ = 0). Rearranging the expression for RA,ef bearing action between the reinforcement lugs and the cover
and making the substitution λ = 0 yields concrete.10 Despite this, a sufficiently accurate picture of the
overall deformation is assumed to be obtained by ignoring
RA, ef − Es As the presence of the reinforcement lugs. This simplification
Ac , ef = (5)
Ec also avoids unnecessary complication of the numerical
model, which would otherwise require several additional
where RA,ef may be determined by a suitable FE analysis. In parameters to define the geometry of the reinforcement lugs.
this section, the simple case of a cylindrical tension member The validity of this assumption will be tested against exper-
is considered to determine values of Ac,ef for a wide variety imental results in following sections. The reinforcement is
of reinforcement ratios and crack spacings. For members therefore idealized as a round bar and the element nodes are
comprising square cross sections, such as those consid- tied along the boundary between the reinforcement and the
ered later in this paper, values of Ac,ef have previously been concrete to satisfy the condition of full strain compatibility.
shown22 to be practically identical to those of cylindrical Both reinforcement and concrete were modeled as linear-
segments having the same reinforcement ratio ρ. elastic isotropic materials and a range of typical values for
The software ABAQUS23 was used to create an axisym- the modular ratio (n = 5 to 8) was considered. A value of Es =
metric FE model of the shaded region in Fig. 3(a), repre- 200 GPa (29,000 ksi) was adopted in the analyses, although
senting a segment of a cylindrical RC tension member the magnitude of both Es and Ec is entirely arbitrary for any
situated between two adjacent primary cracks. To present the given value of n. Poisson’s ratios for steel and concrete were
results of the FE study in a dimensionless form, the geometry taken to be υs = 0.3 and υc = 0.2, respectively. Values of ρ
of the segment is defined in terms of its reinforcement ratio between 0.001 and 0.1 were considered, corresponding to a
ρ = As/Ac and the ratio of the crack spacing to the reinforcement cover dimension in the range of c ≈ 1db to 15db. Values of
bar diameter s/db. The dimensionless form of Eq. (5) is then s/db were taken within the range of 2 to 50. The combina-
tions of n, ρ, and s/db considered herein cover practically all
realistic cases for cracked RC members. Both reinforcement

130 ACI Structural Journal/January 2018


and concrete were modeled with a suitably fine mesh of
eight-node axisymmetric elements (CAX8), the dimension
db taken to be the width of 16 finite elements. Based on the
results of a convergence study, this degree of discretization
was found to represent the best compromise between accu-
racy and computation time. For each combination of n, ρ,
and s/db, a value of ρef was determined according to Eq. (6).
These represented approximately 750 separate FE analyses.
For a very fine spacing of primary cracks (s/db → 0), little
opportunity exists for axial force to be transferred from
the reinforcement to the concrete and the effective area of
concrete approaches zero (ρef → ∞). For very widely spaced
primary cracks (s/db → ∞), the end regions characterized by
non-plane deformation are small compared to the total length
of the segment, and the uncracked behavior is approached
(ρef → ρ). A function that satisfies these criteria is given by Fig. 4—FE results for the case of n = 6.
increasing load and under the influence of shrinkage. These
ρ( s /db )1.2 + A
ρef = (7) are summarized in Table 1.
( s /db )1.2 From each of the studies listed in Table 1, selected tension
stiffening tests were chosen based on the completeness of
where the exponent of 1.2 was found to provide the best the published data, particularly with respect to the load-
overall fit to the entire family of curves produced by the FE deformation curves, material properties, and the spacing of
analyses. The variable A has been determined by regression primary cracks at various load levels. For one study, the
analysis for each combination of n and ρ, and may be taken as elastic modulus of concrete had to be estimated based on the
observed stiffness of the members before cracking. However,
1.13ρ + 0.0458 in most cases, sufficiently detailed data were provided such
A= − 0.048n (8) that the evolution of λ could be determined directly from the
ρ + 0.0255
relevant publications.
Both normal-strength and high-strength concrete members
Figure 4 shows the FE results for the case of n = 6 along- were investigated with mean compressive strengths fcm
side the curves generated by Eq. (7). The error associated between 21.6 and 81.0 MPa (3130 and 11,750 psi). In most
with Eq. (7) is in the order of 2% for the range of n, ρ, and cases, tension members comprised a square cross section
s/db considered herein. with a single centrically embedded ribbed steel reinforce-
ment bar. A range of reinforcement bar diameters (12 to
CHARACTERIZATION OF SCALAR DAMAGE 19.5 mm [0.47 to 0.77 in.]) and reinforcement ratios (0.0099
PARAMETER to 0.0309) was investigated. The tests by Fronteddu26 and
Whereas the influence of primary cracks on the overall Fields and Bischoff27 comprised tension members with
stiffness of an RC tension member may be readily deter- multiple reinforcement bars, for which the tributary concrete
mined by numerical methods (as in the previous section), the cross section surrounding each bar was square.
effect of deteriorating bond on the tension stiffening effect The loading regime is characterized by one of two types:
is governed by a combination of complex, nonlinear effects, purely monotonic loading (M) or a combination of mono-
including10: 1) the gradual formation of a series of closely tonic loading and cycles of loading and unloading (M-C).
spaced internal cracks around the embedded reinforcement; Whereas tests performed with M-C type loading provide a
2) the interaction between the resulting comblike concrete direct measure of the instantaneous stiffness, tests performed
and the reinforcement lugs; 3) localized crushing of the with M type loading must make an adjustment to the apparent
concrete in front of the reinforcement lugs; and at higher axial rigidity to account for any deformation associated with
load levels; and 4) a loss of confinement accompanying the early shrinkage. This is discussed in the following section.
formation of splitting cracks.
Various models are available in the literature to describe Effect of early shrinkage on residual deformation
the deterioration of bond (for example, bond stress-slip For an RC tension member subjected to early shrinkage,
models13-15 and numerical models focused at the meso- an appropriate adjustment must be made to account for the
scale24,25), but the computational demand of many of these shrinkage-induced deformations present in the member
approaches is prohibitive and not conducive to hand calcu- prior to first loading. Figure 5 shows the typical relation-
lations. The scalar damage parameter λ may instead be char- ship between load and average reinforcement strain for an
acterized directly based on the results of tension stiffening RC tension member subjected to an early shrinkage strain of
experiments in the literature. In this section, the results magnitude, εsh. The shrinkage induces an initial compressive
of five tension stiffening studies on axially loaded pris- strain εsm,sh in the reinforcement and shifts the origin of the
matic members are used to describe the evolution of λ with monotonic loading curve from O to O′. For members that

ACI Structural Journal/January 2018 131


Table 1—Summary of tension-stiffening experiments analyzed in this study
Cross section Early shrinkage
Study Loading Member fcm, MPa (psi) n db, mm (in.) dimensions, mm (in.) ρ (×10–6)
UC1 100 x 100* (3.94 x 3.94) 0.0309 –234
Fronteddu26 M-C UC2 33.5 (4860) 6.0 19.5 (0.77) 125 x 125* (4.92 x 4.92) 0.0196 –191
UC5 175 x 175 (6.89 x 6.89)
*
0.0099 –130
NS:15M 41.2 (5980) 6.9 –232
16.0 (0.63) 0.0130
Fields and HS:15M 81.0 (11,750) 4.9 –337
M 125 x 125* (4.92 x 4.92)
Bischoff27 NS:20M 54.9 (7960) 6.9 –261
19.5 (0.77) 0.0196
HS:20M 81.0 (11,750) 5.2 –337
N 35 24.8 (3600) 9.9 †
0
Lee and
M M 35 60.7 (8800) 6.6 †
19 (0.75) 150 x 155 (5.91 x 6.10) 0.0124 0
Kim28
H 35 80.4 (11,660) 6.6 †
0
STN12 12 (0.47) 0.0114 0
21.6 (3130) 8.9
Wu and STN16 16 (0.63) 0.0205 0
M 100 x 100 (3.94 x 3.94)
Gilbert29 STS12 12 (0.47) 0.0114 –249
24.7 (3580) 9.3
STS16 16 (0.63) 0.0205 –249
T1 –222
T2 0
T3 0
Present study M-C T4 35.7 (5180) 7.3 16 (0.63) 100 x 100 (3.94 x 3.94) 0.0205 0
T5 0
T6 0
T7 0
*
For tributary cross section.

Estimated.
Notes: M is monotonic; and M-C is monotonic with cycles of loading and unloading.

are initially uncracked, the age-adjusted effective modulus


method (AEMM)2,30 may be used to determined εsm,sh

1
ε sm , sh = ε sh (9)
1 + ne ρ

where n̄ e is the age-adjusted effective modular ratio and is


given by n̄ e = n(1 + χφ), where χ is an aging coefficient (and
may be taken as approximately 0.8 for problems involving
relaxation2); and φ is the tensile creep coefficient associ-
ated with the period of early shrinkage. For a short member
whose ends represent initial primary cracks, an appropriate
value of ρef may replace ρ in Eq. (9).
The curve OʹVX therefore represents the monotonic enve-
lope for a tension member subjected to early shrinkage, where
V corresponds to the formation of the first primary crack. For
a sufficiently long member (for which the influence of the end
regions may be ignored), the initial branch of the curve OʹV
is defined by the uncracked axial rigidity RA1 = EsAs + EcAc.
Fig. 5—Load-deformation response of RC tension member Alternatively, where the influence of the end regions is to be
subjected to early shrinkage. accounted for, an appropriate value of ρef may be determined
and RA,ef = EsAs + EcAc,ef applies. An assumption of perfect
bond (λ = 0) is reasonable for at least the initial portion of the
loading, although some localized interface damage is likely to
occur even under moderate loading for members whose ends

132 ACI Structural Journal/January 2018


represent initial primary cracks. This is discussed in greater
detail in the following section. The branch VX describes the
gradual formation of primary cracks and deterioration of bond
with increasing load.
As the tension stiffening effect diminishes under
increasing load, so too does the degree of restraint provided
by the embedded reinforcement to the surrounding concrete.
In turn, any residual deformation (that is, the residual
strain under zero load) associated with the period of early
shrinkage must also reduce. For the case of complete bond
deterioration (λ = 1), the reinforcement strain is unaffected
by any shrinkage in the concrete and the residual deforma-
tion must be close to zero.
Suppose that the tension member was subjected to an axial
force corresponding to the point W along the curve OʹVX,
followed by cycles of loading and unloading at this load level
(that is, M-C type loading). The load-deformation response
departs from the monotonic envelope and instead follows
a new path WY. Near point W, this branch of the curve is
linear and its slope represents the effective axial rigidity for
the current state of cracking and bond damage RA,ef = EsAs +
(1 – λ)EcAc,ef. As the tension force in the member approaches
zero, the rough surfaces of the primary cracks meet and are Fig. 6—(a) Testing of member T1; (b) schematic diagram of
unable to fully close. In experiments, this is observed as test setup showing external measurement frame (all dimen-
an apparent increase in axial rigidity and a deviation from sions in mm) and LVDTs (front and rear LVDTs not shown);
the linear part of the loading-unloading curve. The point and (c) end plate detail indicating positions of rods in plan
Yʹ therefore represents the theoretical residual strain due and locking bolts to secure frame to reinforcement bar.
to early shrinkage for a member with smooth crack faces, (Note: 1 mm = 0.0394 in.)
whose magnitude may be taken as2,30 (λ = 0) is assured at the reinforcement-concrete interface.
For the present experimental study, precise measurements
1 of the average reinforcement strain εsm allowed the initial
ε sm
′ , sh = ε sh (10)
ne ρef axial rigidity RA,exp to be accurately calculated. Unlike many
1+
1− λ tension stiffening studies in the literature (which measure
strain at the concrete surface), average reinforcement strains
where ρef = As/Ac,ef is the effective reinforcement ratio for the were directly measured using a stiff external frame attached
current state of cracking. Equation (10), therefore, accounts to the exposed ends of the reinforcement (Fig. 6(a)). Four
for the load-dependent deterioration of restraint provided LVDTs (one at each of the members’ faces) measured the
by the embedded reinforcement to the now cracked and total elongation of the reinforcement bars and an appro-
damaged concrete. priate adjustment was made to account for the additional
For purely monotonic loading (M), the apparent axial elongation of the bars’ exposed ends (Fig. 6(b) and 6(c)).
rigidity represented by the secant OW would overesti- The members in this study were tested with a 1 MN (225 kip)
mate RA,ef in cases where early shrinkage occurs and is capacity universal testing machine using a combination of
not measured. Rather, RA,ef is defined by the instantaneous monotonic loading and cycles of loading and unloading
response YʹW. For tests involving monotonic loading and (M-C type loading). Members T1 and T2 were initially
cycles of loading and unloading (M-C), this distinction subjected to a monotonic load of 35 kN (7.9 kip) followed
becomes apparent and RA,ef may be determined from the by five cycles of loading and unloading at each of several
slope of YʹW directly. Thus, the adjustment in Eq. (10) allows increasing load levels of 5 kN (1.1 kip) increments until
the effective axial rigidity of cracked RC tension members a maximum load of 100 kN (22.5 kip), close to the yield
to be directly compared, regardless of the type of loading point of the reinforcement. Members T3 to T7, which form
performed or of any early shrinkage that has occurred prior part of an ongoing study, were subjected to a monotonically
to first loading. In any case, after the formation of a stabi- increasing load of 50 kN (11.2 kip) followed by five cycles
lized crack pattern, much of the initial restraint is lost and the of loading and unloading at that load level only. For each of
effect of early shrinkage on residual deformations is small. the seven members tested, RA,exp was determined by fitting
a straight line of best fit to the initial portion of the load
Initial behavior (λ = 0) versus average reinforcement strain graph prior to cracking
The initial response of the RC tension members to mono- for loads between approximately 0 to 5 kN (0 to 1.1 kip).
tonic loading presents a special opportunity for which the The members tested in the present study had nominal
validity of Eq. (7) may be tested. This is because for a spec- dimensions of 100 x 100 x 900 mm (3.93 x 3.93 x 35.43 in.)
imen not previously loaded, a condition of perfect bond and were reinforced with a single 16 mm (0.63 in.) diameter

ACI Structural Journal/January 2018 133


Fig. 8—Full load-deformation response of member T1.
Primary cracks indicated by numbers 1 to 5; formation of
first splitting cracks indicated by S.

Fig. 7—Initial load-deformation response of members T1 to Table 2—Loads (kN) and positions (mm) of
T7. Early shrinkage for member T1 has been ignored here. primary cracks for present experimental study
Member Crack 1 Crack 2 Crack 3 Crack 4 Crack 5
steel reinforcement bar (ρ = 0.0205). The initial crack
spacing may be taken as the length of the prism and, there- T1 23.8; 218 25.9; 656 32.2; 356 35.0; 519 39.0; 123
fore, s/db = 56.25. Values of Es and Ec were found to be T2 33.3; 505 39.6; 303 45.0; 664 49.6; 142 —
203 and 27.7 GPa (29,400 and 4040 ksi), respectively (n = 7.3).
From Eq. (7), the effective reinforcement ratio is calculated T3 41.5; 676 46.4; 365 — — —
as ρef = 0.0296. Assuming perfect bond (λ = 0), the theoret- T4 42.0; 230 46.0; 407 47.2; 674 — —
ical initial effective rigidity for these members is RA,ef = T5 41.8; 581 44.5; 322 — — —
229 MN (51,500 kip). Even for such a wide spacing of
primary “cracks”, the influence of the non-uniform deforma- T6 40.1; 404 47.2; 607 — — —
tion in the end regions is significant; for comparison, the T7 43.9; 628 46.0; 273 — — —
axial rigidity of the uncracked member is RA1 = 312 MN
Notes: Crack positions are measured from the top of each member; 1 kN = 0.225 kip;
(70,100 kip), some 36% greater. 1 mm = 0.0394 in.
Figure 7 shows part of the load-deformation curves for
each of the seven members tested in this study. The initial Figure 8 presents an overview of the total load-deformation
portions of the curves closely match the effective axial response of member T1 tested in the present experimental
rigidity of RA,ef = 229 MN (51,500 kip) predicted by Eq. (7), study. A summary of the formation of primary cracks for the
whereas the uncracked rigidity RA1 = 312 MN (70,100 kip) present study is given in Table 2. For clarity, only the
obviously overestimates the initial response. The average ascending portion of the fifth load cycle at each load level is
initial rigidity for members T1 to T7 is measured to be 233 plotted. In addition to the primary cracks, splitting cracks
MN (52,300 kip), just 2% greater than that suggested by were observed to form along each of the seven tension
Eq. (7). This result confirms the previous assumption that members tested in the present study (T1 to T7). These were
the effect of the reinforcement lugs on the overall deforma- first observed at loads of approximately 50 kN (11.2 kip),
tion is negligible for the case of perfect bond. and for members T1 and T2, the splitting cracks increased in
length and width with increasing loads up to 100 kN
Deterioration of bond (λ > 0) (22.5 kip). After the formation of a stabilized crack pattern,
In Fig. 7, for a tension force less than approximately 10 kN the stiffness of the load-deformation response is seen to
(2.2 kip), the initial slopes of the loading curves closely match gradually reduce with each load increment. At a load of
the theoretical rigidity for the case of λ = 0. Soon thereafter, a 100 kN (22.5 kip), which is close to the yield point of the
softening effect is observed and the curves gradually begin to reinforcement, the response of the cracked tension member
deviate from the case of perfect bond. Because the spacing of is almost indistinguishable from that of the bare reinforce-
the primary (end) “cracks” is constant before the formation of ment, indicating a near-complete deterioration of bond. The
the first primary crack, any deviation from the initial rigidity apparent increase in axial rigidity due to the closing cracks
must be a result of bond deterioration (that is, λ > 0). At higher (as highlighted in Fig. 5) is evident in the response of T1.
loads, this effect is observed to continue until the tension stiff-
ening effect is almost entirely exhausted.

134 ACI Structural Journal/January 2018


Effect of shrinkage on bond deterioration
As concrete shrinks, it is restrained by embedded
reinforcement and as a result develops tensile stress. This
restraint is provided by bond stresses that develop at the
reinforcement-concrete interface near the primary cracks. For
a member subjected to early shrinkage, therefore, a certain
degree of bond is demanded even before external loads are
applied. Thus, for a given load level, a member that has been
subjected to early shrinkage demands a greater degree of bond
compared to a member that is free of shrinkage. In turn, a
greater degree of bond deterioration would be expected for the
member subjected to early shrinkage.
One way to account for this effect is to consider the shrink-
age-induced tensile stress in the concrete to be the result
of a fictitious mechanical load applied to the ends of the
reinforcement. In this way, the additional deterioration of
bond may be determined by superimposing the fictitious
shrinkage load with existing mechanical loads in the deter-
mination of λ. This is the basic methodology of Murray et al.6
Further explanation of this approach may be found in that
study. The magnitude of this fictitious axial force is determined
using the AEMM,2,6,30 and may be expressed in terms of the
resulting strain increment in the bare reinforcement εs,sh
Fig. 9—Evolution of bond damage with increasing load level.
1 + nρef sm is the average crack spacing. In other words, for deter-
ε s , sh = − ε sh (11)
ne ρef mining values of λ from the load-deformation responses of
1+ the members, bond damage may be deemed to occur within
1− λ
a maximum distance of approximately 10db either side of
where the shrinkage strain εsh is negative. In Eq. (11), λ a primary crack. Elsewhere, a condition of perfect bond is
represents the degree of bond deterioration that exists prior assumed. Importantly, this adjustment is only necessary for
to the development of the shrinkage strain. Equation (11) the initial stages of loading; for each of the tension members
therefore correctly accounts for the load-dependent deterio- analyzed in this study, the primary crack spacing at the stabi-
ration of restraint provided by the embedded reinforcement. lized crack stage satisfied the requirement of sm/db < 20.
Where limited early shrinkage occurs before first loading, a
value of λ = 0 may be safely assumed. Evolution of λ
Each of the tension-stiffening tests summarized in Table 1
Adjustment for wide crack spacing was analyzed to determine values of λ over a wide range of
For a segment located between widely spaced primary load levels. For each test, Eq. (4) was used to yield approx-
cracks, it is difficult to characterize the state of interface imately 10 values of λ based on the observed load-deformation
damage of the entire segment with a single value of λ. Near response and the precise spacing of primary cracks, as
the primary cracks, where reinforcement slip is greatest, reported by the respective authors. For tests conducted with
extensive internal cracking and splitting of the cover concrete purely monotonic loading (M), an adjustment was made in
may occur, while a state of perfect bond may exist in regions accordance with Eq. (10) to account for the influence of any
remote from the cracks. This is particularly so during initial residual deformations associated with early shrinkage on the
loading, when the average primary crack spacing is rela- axial rigidity of the member. Where cycles of loading and
tively large yet interface damage demonstrably begins to unloading were also performed (M-C), the instantaneous
occur even for modest loading (Fig. 7). To account for this, response of the member was used directly. For members
the size of the region near a primary crack that is considered subjected to early shrinkage, a fictitious strain εs,sh (Eq. (11))
to exhibit bond deterioration may be limited. was added to the bare reinforcement strain εs2 to accompany
Because the extent of bond damage (as characterized by each value of λ. A plot of λ versus the adjusted reinforcement
λ) should ideally be independent of the spacing of primary strain (εs2 + εs,sh) is shown in Fig. 9.
cracks (as it is intended to be), it follows that values of λ Overall, there appears to be little discernible influence
determined immediately before and after the formation of on the evolution of λ of concrete strength grade, reinforce-
a single primary crack should be approximately equal. In ment ratio, or reinforcement bar diameter (albeit within the
other words, at a constant load level, the formation of a new limited ranges of these parameters considered in this paper).
primary crack should be able to account for the entire reduc- In many cases, variation of λ at a given load level is greater
tion in stiffness that is observed. For the tension members among nominally identical members than among members
analyzed in this paper, this was found to be the case, from different studies. As demonstrated previously for the
provided that sm/db is less than approximately 20, where present study (Fig. 7), initiation of bond damage is seen in

ACI Structural Journal/January 2018 135


Fig. 9 to occur soon after first loading. Not surprisingly, for
an adjusted reinforcement strain (εs2 + εs,sh) less than approx-
imately 300 × 10–6, values of λ were found to be approx-
imately zero. Thereafter, the degree of bond deterioration
increases rapidly and then gradually approaches a value of
λ = 1 near the yield point of the reinforcement (εs2 ≈ 2500 ×
10–6). A curve of best fit is given by Eq. (12). Approximately
90% of the values of λ calculated from the experimental
results fall within ±0.1 of this curve

λ = 0 for εs2 + εs,sh < 300 × 10–6 (12a)

λ = 1 – exp[–1600(εs2 + εs,sh – 300 × 10–6)]


for εs2 + εs,sh ≥ 300 × 10–6 (12b)

VALIDATION OF MODEL
The tension-stiffening model proposed in this paper may
Fig. 10—Simulation of three tension-stiffening tests by
be used in conjunction with any deterministic or stochastic
Abrishami and Mitchell.31
model of primary crack formation. Alternatively, the general
approach outlined in this paper may be used to develop a FURTHER RESEARCH
new model of primary crack formation based on consider- Further experimental investigation would be valuable to
ations of highly non-uniform strain regions near existing characterize the influence of a wider range of concrete
primary cracks. In this section, the proposed model is used to strength grades, reinforcement ratios, and bar diameters on
simulate the load-deformation responses of three RC tension the evolution of the bond damage parameter λ. The effects of
members tested by Abrishami and Mitchell.31 The spacing of long-term shrinkage and sustained loading on time-dependent
primary cracks at the stabilized cracking stage (as reported changes in tension stiffening form the basis of ongoing
by the authors) is ignored in this analysis. Instead, a simple experimental studies by the authors.
code-based model of primary crack formation is adopted.32
The first primary crack is assumed to occur at a load of Ncr CONCLUSIONS
when the average concrete stress in a region of plane defor- Based on the findings of this study, the following conclu-
mation reaches the effective tensile strength of the concrete sions can be made:
(that is, accounting for the influence of early shrinkage). 1. The separate influences of primary crack formation and
Prior to the formation of the first primary crack, the tension bond deterioration on the overall load-deformation response
member is treated as a single segment located between end of RC tension members were modeled based on the results
“cracks,” and the initial response is calculated according to of a FE study and experimental tension stiffening studies in
Eq. (7) assuming that s = L, where L is the total member the literature, respectively.
length. An assumption of uncracked behavior during the 2. For the case of perfect bond (that is, where no slip
early loading period would lead to a significant overestima- occurs along the reinforcement-concrete interface), the pres-
tion of initial rigidity compared to the experimental response ence of primary cracks alone represents a major mechanism
(refer to Fig. 7). of the tension stiffening effect. This is attributed to the exis-
At a load greater than or equal to 1.5Ncr, a stabilized crack tence of highly non-uniform strain regions in the concrete
pattern is assumed to have formed, for which the average surrounding the embedded reinforcement near the primary
primary crack spacing (in mm) may be taken as cracks. FE modeling for the case of perfect bond is shown
to accurately predict the initial load-deformation response of
db RC tension members tested in a recent study by the authors.
sm = 50 + 0.25k1k2 (13)
ρ 3. For members whose end regions represent initial
primary cracks, deterioration of bond is found to begin soon
after the first application of load. For other types of members,
where, for a tension member with deformed reinforcement
whose end regions are not subjected to significant tension
bars, k1 = 0.8 and k2 = 1.0.32 The number of cracks is assumed
forces, deterioration of bond begins immediately after the
to be proportional to the magnitude of load between Ncr and
formation of the first primary crack. This bond deterioration
1.5Ncr (that is, sm is taken to be inversely proportional to the
is attributed to the formation of internal cone-shaped cracks
load within this range). Overall, good agreement is achieved
around the embedded reinforcement, and at higher loads to
between the proposed model and the test results (Fig. 10).
the effects of crushing of concrete in front of the reinforce-
Particularly, the current model improves markedly upon that
ment lugs and splitting of the cover concrete.
proposed in the study by Abrishami and Mitchell.31
4. Bond deterioration is characterized by the introduction
of a scalar damage parameter λ that depends on the magni-
tude of applied loading and shrinkage. The evolution of λ
with increasing load appears to be insensitive to factors such

136 ACI Structural Journal/January 2018


as concrete strength grade, reinforcement ratio, and rein- Engineering, ASCE, V. 118, No. 8, 1992, pp. 2118-2132. doi: 10.1061/
(ASCE)0733-9445(1992)118:8(2118)
forcement bar diameter (within the ranges of these parame- 13. Fédération Internationale du Béton, “Model Code for Concrete
ters considered in this paper). Structures 2010,” Ernst & Sohn, Berlin, Germany, 2013.
14. Muhamad, R.; Ali, M. S. M.; Oehlers, D. J.; and Griffith, M., “The
Tension Stiffening Mechanism in Reinforced Concrete Prisms,” Advances
AUTHOR BIOS in Structural Engineering, V. 15, No. 12, 2012, pp. 2053-2069. doi:
Angus Murray is a PhD Candidate in the School of Civil and Environmental
10.1260/1369-4332.15.12.2053
Engineering at the University of New South Wales, Kensington, Australia.
15. Yankelevsky, D. Z.; Jabareen, M.; and Abutbul, A. D., “One-Dimen-
His research interests include serviceability of concrete structures.
sional Analysis of Tension Stiffening in Reinforced Concrete with Discrete
Cracks,” Engineering Structures, V. 30, No. 1, 2008, pp. 206-217. doi:
ACI member Raymond Ian Gilbert is Emeritus Professor and Deputy
10.1016/j.engstruct.2007.03.013
Director of the Centre for Infrastructure Engineering and Safety in the
16. Saint-Venant, A. J. C. B., “Mémoire sur la torsion des prismes,”
School of Civil and Environmental Engineering at the University of New
Mémoires des Savants Étrangers, V. 14, 1856.
South Wales. His research interests include serviceability of reinforced
17. Watstein, D., and Mathey, R. G., “Width of Cracks in Concrete at the
and prestressed concrete structures. He is a member of ACI Committees
Surface of Reinforcing Steel Evaluated by Means of Tensile Bond Speci-
209, Creep and Shrinkage in Concrete; and 318, Structural Concrete
mens,” ACI Journal Proceedings, V. 56, No. 7, July 1959, pp. 47-56.
Building Code.
18. Broms, B. B., “Crack Width and Crack Spacing in Reinforced
Concrete Members,” ACI Journal Proceedings, V. 62, No. 10, Oct. 1965,
Arnaud Castel is an Associate Professor at the School of Civil and Envi-
pp. 1237-1255.
ronmental Engineering at the University of New South Wales. His research
19. Yannopoulos, P. J., “Variation of Concrete Crack Widths through the
interests include durability of construction materials, low-carbon concrete,
Concrete Cover to Reinforcement,” Magazine of Concrete Research, V. 41,
and corrosion effects in concrete structures.
No. 147, 1989, pp. 63-68. doi: 10.1680/macr.1989.41.147.63
20. Gergely, P., and Lutz, L. A., “Maximum Crack Width in Rein-
ACKNOWLEDGMENTS forced Concrete Flexural Members,” Causes, Mechanism, and Control of
This research was supported by the Australian Research Council through Cracking in Concrete, SP-20, American Concrete Institute, Farmington
ARC Discovery Projects DP110103028 and DP14010052. The authors Hills, MI, 1968, pp. 1-17.
wish to thank A. Macken for his assistance with the experiments. 21. Frosch, R. J., “Another Look at Cracking and Crack Control in
Reinforced Concrete,” ACI Structural Journal, V. 96, No. 3, May-June 1999,
pp. 437-442.
REFERENCES 22. Murray, A.; Gilbert, R. I.; and Castel, A., “Study of Disturbed Strain
1. Marti, P.; Alvarez, M.; Kaufmann, W.; and Sigrist, V., “Tension Chord Regions for Serviceability Analysis of Cracked RC Flexural Members,”
Model for Structural Concrete,” Structural Engineering International, V. 8, 24th Australasian Conference on the Mechanics of Structures and Mate-
No. 4, 1998, pp. 287-298. doi: 10.2749/101686698780488875 rials, H. Hong, H. and C. Zhang, eds., 2017, pp. 97-102.
2. Gilbert, R. I., and Ranzi, G., Time-Dependent Behaviour of Concrete 23. Dassault Systèmes Simulia Corp, ABAQUS/Standard User’s Manual
Structures, Spon Press, London, UK, 2011. V. 6.12, Providence, RI, 2012.
3. Kwak, H. G., and Song, J. Y., “Cracking Analysis of RC Members 24. Michou, A.; Hilaire, A.; Benboudjema, F.; Nahas, G.; Wyniecki, P.;
Using Polynomial Strain Distribution Function,” Engineering Structures, and Berthaud, Y., “Reinforcement-Concrete Bond Behavior: Experimenta-
V. 24, No. 4, 2002, pp. 455-468. doi: 10.1016/S0141-0296(01)00112-2 tion in Drying Conditions and Meso-Scale Modeling,” Engineering Struc-
4. Castel, A.; Gilbert, R. I.; and Ranzi, G., “Instantaneous Stiffness of tures, V. 101, 2015, pp. 570-582. doi: 10.1016/j.engstruct.2015.07.028
Cracked Reinforced Concrete Including Steel-Concrete Interface Damage 25. Hayashi, D., and Nagai, K., “Meso-Scale Approach to Investigate
and Long-Term Effects,” Journal of Structural Engineering, ASCE, V. 140, the Bond Performance of Anchorage in Reinforced Concrete,” 9th Interna-
No. 6, 2014, p. 04014021 doi: 10.1061/(ASCE)ST.1943-541X.0000954 tional Conference on Fracture Mechanics of Concrete and Concrete Struc-
5. Xu, T.; Castel, A.; Gilbert, R. I.; and Murray, A., “Modeling the tures, V. Saouma, J. Bolander, and E. Landis, eds., University of California,
Tensile Steel Reinforcement Strain in RC-Beams Subjected to Cycles of Berkeley, Berkeley, CA, 2016. doi: 10.21012/FC9.123
Loading and Unloading,” Engineering Structures, V. 126, 2016, pp. 92-105. 26. Fronteddu, L. F., “Response of Reinforced Concrete to Reverse Cyclic
doi: 10.1016/j.engstruct.2016.07.043 Loading,” master’s thesis, University of British Columbia, Vancouver, BC,
6. Murray, A.; Castel, A.; Gilbert, R. I.; and Chang, Z.-T., “Time- Canada, 1992.
Dependent Changes in the Instantaneous Stiffness of Reinforced Concrete 27. Fields, K., and Bischoff, P. H., “Tension Stiffening and Cracking
Beams,” Engineering Structures, V. 126, 2016, pp. 641-651. doi: 10.1016/j. of High-Strength Reinforced Concrete Tension Members,” ACI Structural
engstruct.2016.08.025 Journal, V. 101, No. 4, July-Aug. 2004, pp. 447-456.
7. Scott, R. H., and Gill, P. A. T., “Short Term Distributions of Strain and 28. Lee, G.-Y., and Kim, W., “Cracking and Tension Stiffening Behavior
Bond Stress Along Tension Reinforcement,” The Structural Engineer, of High-Strength Concrete Tension Members Subjected to Axial Load,”
V. 65B, No. 2, 1987, pp. 39-43. Advances in Structural Engineering, V. 11, No. 5, 2008, pp. 127-137.
8. Kenel, A.; Nellen, P.; Frank, A.; and Marti, P., “Reinforcing Steel 29. Wu, H. Q., and Gilbert, R. I., “An Experimental Study of Tension
Strains Measured by Bragg Grating Sensors,” Journal of Materials in Stiffening in Reinforced Concrete Tension Members under Short-Term and
Civil Engineering, ASCE, V. 17, No. 4, 2005, pp. 423-431. doi: 10.1061/ Long-Term Service Loads,” UNICIV Report No. R-449, University of New
(ASCE)0899-1561(2005)17:4(423) South Wales, Kensington, NSW, Australia, 2008, 31 pp.
9. Goto, Y., “Cracks Formed in Concrete around Deformed Tension 30. Trost, H., “Auswirkungen des Superpositionsprinzips auf Kriech-
Bars,” ACI Journal Proceedings, V. 68, No. 4, Apr. 1971, pp. 244-251. und Relaxations-Probleme bei Beton und Spannbeton,” Beton- und Stahl-
10. Fédération Internationale du Béton, Bulletin 10: Bond of reinforce- betonbau, V. 62, No. 10, 1967, pp. 230-238., 261-269.
ment in concrete, fib, Lausanne, Switzerland, 2000. 31. Abrishami, H. H., and Mitchell, D., “Influence of Splitting Cracks on
11. Somayaji, S., and Shah, S. P., “Bond Stress versus Slip Relationship Tension Stiffening,” ACI Structural Journal, V. 93, No. 6, Nov.-Dec. 1996,
and Cracking Response of Tension Members,” ACI Journal Proceedings, pp. 703-710.
V. 78, No. 3, May-June 1981, pp. 217-225. 32. British Standards Institution, “Eurocode 2: Design of Concrete Struc-
12. Chan, H. C.; Cheung, Y. K.; and Huang, Y. P., “Crack Analysis tures – Part 1: General Rules and Rules for Buildings – DD ENV 1992-1-
of Reinforced Concrete Tension Members,” Journal of Structural 1:1992,” European Committee for Standardization, Brussels, Belgium, 1992.

ACI Structural Journal/January 2018 137


CALL FOR ACTION
ACI Invites You To...
Do you have EXPERTISE in any of these areas?
• BIM
• Chimneys
• Circular Concrete Structures Prestressed by Wrapping
with Wire and Strand
• Circular Concrete Structures Prestressed with
Circumferential Tendons
• Concrete Properties
• Demolition
• Deterioration of Concrete in Hydraulic Structures
• Electronic Data Exchange
• Insulating Concrete Forms, Design, and Construction
• Nuclear Reactors, Concrete Components
• Pedestal Water Towers
• Pipe, Cast-in-Place
• Strengthening of Concrete Members
• Sustainability

Then become a REVIEWER for the


ACI Structural Journal or the ACI Materials Journal.
How to become a Reviewer:
1. Go to: http://mc.manuscriptcentral.com/aci;
2. Click on “Create Account” in the upper right-hand corner; and
3. Enter your E-mail/Name, Address, User ID and Password, and
Area(s) of Expertise.

Did you know that the database for MANUSCRIPT


CENTRAL, our manuscript submission program,
is separate from the ACI membership database?
How to update your user account:
1. Go to http://mc.manuscriptcentral.com/aci;
2. Log in with your current User ID & Password; and
3. Update your E-mail/Name, Address, User ID and Password,
and Area(s) of Expertise.

QUESTIONS?
E-mail any questions to Journals.Manuscripts@concrete.org.
ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 115-S11

Punching Shear Strength of Slabs and Influence of Low


Reinforcement Ratio
by Susanto Teng, Khatthanam Chanthabouala, Darren T. Y. Lim, and Rhahmadatul Hidayat
Presented in this paper are the punching shear tests of 12 high- ural failure under punching load may occur ahead of the
strength concrete slabs (fc′ > 100 MPa [14,500 psi]) with various normal punching shear failure.
flexural reinforcement ratios from 0.28 to 1.43% and supported on One typical way to treat this lightly reinforced slab case
columns having various aspect ratios of 1 x 1, 1 x 3, and 1 x 5. The in practice is to calculate the punching strength of the slab
resulting test data were also used to verify the suitability of the
using the ACI punching shear equations and compare it with
ACI 318 and Eurocode 2 punching shear equations, especially for
the failure load (ultimate shear force) that causes flexural
slabs with low reinforcement ratios.
Other influencing factors such as concrete strength and size effect failure of the slab as calculated using the yield line theory.
were also addressed. The authors’ new experimental results will be The lower of these two failure loads will be the punching
combined with 355 existing published data and together they will shear capacity of the slab. However, it would be much
be used to evaluate the accuracies and safety of the ACI 318 and easier if Code method can treat cases of low reinforcement
Eurocode 2 methods for punching shear, as well as some methods ratio directly.
proposed by other researchers. A new general method and a simpli- To address the issues mentioned previously, the authors
fied method are also proposed; they are shown to be accurate present an experimental program involving the testing of 12
and reliable. high-strength concrete slabs having various flexural rein-
forcement ratios and various column aspect (rectangularity)
Keywords: building codes; column aspect ratio; concrete slabs; flexural
failure; high-strength concrete; punching shear; reinforcement ratio; shear ratios subjected to concentric punching loads. The authors
failure; size effect. have previously tested similar slabs made of normal-strength
concrete to investigate the influence of column aspect ratio
INTRODUCTION as well as opening.3 Those earlier tests will be useful in
Punching shear failure is one of the critical failure modes current research. In addition, the authors’ new experimental
that can happen in a flat-plate floor system. Some of the results will also be combined with 355 existing published
important design parameters that can influence the punching data and together they will be used to evaluate the accura-
shear strength of concrete slabs include concrete strength, cies of the ACI 318-14 and Eurocode 2 methods, as well as
amount of flexural reinforcement, column rectangularity the methods proposed by other researchers.4-6 A new general
(aspect) ratio, and size effect. method and a simplified method are also proposed; they are
One of the parameters that has not been sufficiently inves- shown to be accurate and reliable.
tigated in the past is the influence of the amount of flex-
ural reinforcement ratio ρ (=As/bd), especially the low rein- RESEARCH SIGNIFICANCE
forcement ratio. Floor slabs with low reinforcement ratios A new way to treat punching shear strength of concrete
are frequently encountered in the design of flat-plate floors slabs is presented herein. The influence of low flexural
for lightly loaded buildings, such as apartment and condo- reinforcement ratio is considered rationally, resulting in
minium buildings or office buildings. The typical values of two proposed methods: a proposed general method and a
the required flexural reinforcement ratios ρ (=As/bd) in the proposed simplified method. The new proposed methods
column strips of the flat-plate floors in those buildings can were verified for accuracy using the authors’ new experi-
vary from approximately 0.6 to 0.8%. It has been known that ment, which is also presented in this paper, as well as 355
when a floor slab is provided with a low flexural reinforce- slab data from literature. The proposed methods were also
ment ratio (ρ of approximately 0.7% or less), the slab may thoroughly compared with existing methods proposed by
not be able to attain its punching shear capacity as calcu- other researchers as well as with the methods of ACI 318
lated using the ACI 318-141 equations for punching shear. and Eurocode 2.
That is because it may fail at a lower load due to widespread
yielding of the flexural reinforcement. Essentially, that is a EXPERIMENTAL PROGRAM
flexural failure under punching shear load. A further increase The slab specimens in this experimental program were
in loading beyond the peak load will only increase slab rota- designed to cover a low range of reinforcement ratios as
tion or deflection before it reaches the final failure in what well as some of the higher ranges of reinforcement ratios to
looks like a punching failure—albeit a ductile punching ACI Structural Journal, V. 115, No. 1, January 2018.
failure. So far, the ACI 318 or the Eurocode 22 (EC2) equa- MS No. S-2016-432, doi: 10.14359/51701089, was received December 8, 2016, and
reviewed under Institute publication policies. Copyright © 2018, American Concrete
tions for punching shear have not been sufficiently verified Institute. All rights reserved, including the making of copies unless permission is
for cases of slabs with low reinforcement ratios where flex- obtained from the copyright proprietors. Pertinent discussion including author’s
closure, if any, will be published ten months from this journal’s date if the discussion
is received within four months of the paper’s print publication.

ACI Structural Journal/January 2018 139


Fig. 1—(a) General dimensions and loading positions; and (b) reinforcement details.
provide a more complete set of data. The dimensions of the important properties of the specimens. The overall dimen-
specimens represent an actual model of the column zone of sions (L1 x L2 x h) of the specimens were 2.2 x 2.2 x 0.15 m
flat-plate slabs having column-to-column spans of approxi- (87 x 87 x 5.9 in.) for the S11 and S13 series, and 2.7 x 2.2 x
mately 5.0 to 6.0 m (16.4 to 19.7 ft). 0.15 m (106 x 87 x 5.9 in.) for the S15 series. The cross-
sectional dimensions of the column stubs were 200 x
Specimen details and material properties 200 mm (7.9 x 7.9 in.) for the S11 series, 600 x 200 mm
The 12 slab specimens tested in this experiment were cate- (23.7 x 7.9 in.) for the S13 series, and 1000 x 200 mm (39.5 x
gorized into three main groups: S11 series, S13 series, and 7.9 in.) for the S15 series. The height of the column stub was
S15 series to represent their corresponding column aspect 200 mm (7.9 in.) for all specimens. In addition to the main
ratios of 1 x 1, 1 x 3, and 1 x 5, respectively. Figure 1(a) top reinforcement, all the specimens were provided with
shows the general dimensions of the slab specimens and bottom reinforcement in the form of 10 mm (3/8 in.) diam-
their loading positions while Fig. 1(b) shows the reinforce- eter bars that were distributed at 260 mm (10.2 in.) spacing,
ment details of the 12 slab specimens. Table 1 summarizes denoted as T10@260 mm (No. 3@10.2 in.) in Fig. 1(b).

140 ACI Structural Journal/January 2018


Table 1—Properties of slab specimens
fy, MPa Top reinforcement bar size @
No. Slab ID Dimensions, m (in.) Column size, m (in.) d, mm (in.) fc′, MPa (ksi) (ksi) ρ, % spacing, mm (in.)
(1) (2) (3) (4) (5) (6) (7) (8) (9) (10)
1 S11-028 120 (4.7) 459 (66.5) 0.28 T10@ 260 (No. 3 @ 10.2)

S11 2 S11-050 2.2 x 2.2 x 0.15 0.2 x 0.2 117 (4.6) 537 (77.9) 0.50 T13@ 235 (No. 4 @ 9.2)
112.0 (16.2)
Series 3 S11-090 (87 x 87 x 5.9) (7.9 x 7.9) 117 (4.6) 537 (77.9) 0.90 T13@ 118 (No. 4 @ 4.6)
4 S11-139 114 (4.5) 501 (72.6) 1.39 T16@ 118 (No. 5 @ 4.6)
5 S13-028 120 (4.7) 459 (66.5) 0.28 T10@ 260 (No. 3 @ 10.2)

S13 6 S13-050 2.2 x 2.2 x 0.15 0.2 x 0.6 117 (4.6) 537 (77.9) 0.50 T13@ 235 (No. 4 @ 9.2)
114.0 (16.5)
Series 7 S13-090 (87 x 87 x 5.9) (7.9 x 23.7) 117 (4.6) 537 (77.9) 0.90 T13@ 118 (No. 4 @ 4.6)
8 S13-143 114 (4.5) 501 (72.6) 1.43 T16@ 118 (No. 5 @ 4.6)
9 S15-028 120 (4.7) 459 (66.5) 0.28 T10@ 260 (No. 3 @ 10.2)

S15 10 S15-050 2.7 x 2.2 x 0.15 0.2 x 1.0 117 (4.6) 537 (77.9) 0.50 T13@ 235 (No. 4 @ 9.2)
97.0 (14)
Series 11 S15-090 (106.4 x 87 x 5.9) (7.9 x 39.5) 117 (4.6) 537 (77.9) 0.90 T13@ 118 (No. 4 @ 4.6)
12 S15-143 114 (4.5) 501 (72.6) 1.43 T16@ 118 (No. 5 @ 4.6)
Notes: Concrete cover = 20 mm (0.8 in.); maximum aggregate size = 20 mm (0.8 in.); same reinforcement is provided in both directions; T10 @ 260 mm = 10 mm bars at 260 mm
spacing or No. 3 bar (3/8 in.) at 10.2 in. spacing; d is average effective depth fc′ is cylinder compressive strength of concrete; fy is yield strength of flexural reinforcement; ρ is average
reinforcement ratio (ρx + ρy)/2; ρx = 100Asx/(Lx × dx).

Fig. 2—(a) Typical test setup; and (b) Specimen S13-028 during testing.
The specimen notation represents the main properties of through four hydraulic jacks that were secured onto the
the slab specimens. For example, Specimen S13-143 indi- laboratory strong floor. Each hydraulic jack would apply
cates a slab specimen with a column aspect ratio of 1 x 3 the loading by pulling down the steel rod, which transferred
(β = 3) and flexural reinforcement ratio ρ of 1.43%. the pulldown force to the spreader beams and then onto
The strengths of the concrete used in all the speci- the loading plates (points) on the slab. The actual positions
mens were approximately 100 MPa (14,500 psi) (refer of the spreader beams and loading points (Fig. 1(a)) were
to Column 7 of Table 1). The maximum aggregate size is determined using a finite element software such that the
20 mm (0.8 in.). Some of the advantages of using high- distributions of stresses near the column zone were close to
strength concrete compared to normal-strength concrete those stress distributions in the same slab when loaded under
include an increase in cracking load of the slab and a reduc- uniform loading.
tion in deflection at service load level due to higher tensile Strain gauges were installed on some of the top rein-
strength and higher elastic modulus of higher-strength forcing bars in both directions and linear variable differen-
concrete. Higher concrete strength also leads to higher dura- tial transformers (LVDTs) were placed below the slab along
bility in adverse environments. the column center lines to measure vertical deflections at
every load increment.
Instrumentation and loading procedure Each specimen was loaded at 20 kN (4.5 kip) load incre-
A typical test setup and a photograph of a slab during ment or approximately 5 kN (1.12 kip) increment for each
testing are shown in Fig. 2. Each specimen was placed on jack. At every load increment, readings of vertical displace-
a steel support block and then vertically loaded downward ments from LVDTs and steel strains were recorded all the

ACI Structural Journal/January 2018 141


Table 2—Comparison of design methods with current test results
Vexp/Vcalc
Proposed Proposed
c2 x c1, mm d, mm Vexp*, kN Failure ρfs‡, ρ/ρfs§, ACI Eurocode Peiris- General Simplified
Slab ID (in.) (in.) ρ, % (kip) mode† % (4)/(7) 318-14 2 Ghali CSCT Method Method
(1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) (12) (13) (14)
S11-028 120 (4.7) 0.28 280 (62.9) F 0.96 0.29 (F) 0.65 0.96 1.92 1.12 1.19 1.13
S11-050 200 x 200 117 (4.6) 0.50 394 (88.6) F 0.98 0.51 (F) 0.95 1.16 1.58 1.21 1.29 1.22
S11-090 (7.9 x 7.9) 117 (4.6) 0.90 440 (98.9) P 0.78 1.16 (P) 1.06 1.06 1.06 0.99 1.06 1.06
S11-139 114 (4.5) 1.39 454 (102.1) P 0.88 1.58 (P) 1.14 0.99 1.14 0.89 0.98 0.98
S13-028 120 (4.7) 0.28 308 (69.2) F 0.98 0.27 (F) 0.53 0.79 2.18 0.99 1.07 1.01
S13-050 200 x 600 117 (4.6) 0.50 418 (94.0) F 0.80 0.63 (F) 0.74 0.91 1.44 0.97 1.06 1.04
S13-090 (7.9 x 23.7) 117 (4.6) 0.90 558 (125.4) P 0.80 1.13 (P) 0.99 1.00 1.08 0.96 1.08 1.08
S13-143 114 (4.5) 1.43 718 (161.4) P 0.91 1.57 (P) 1.32 1.14 1.32 1.05 1.23 1.23
S15-028 120 (4.7) 0.28 322 (74.4) F 1.07 0.25 (F) 0.49 0.68 2.42 0.97 1.04 0.97
S15-050 200 x 1000 117 (4.6) 0.50 458 (103.0) F 0.85 0.58 (F) 0.70 0.79 1.59 0.95 1.04 1.00
S15-090 (7.9 x 39.5) 117 (4.6) 0.90 658 (147.9) P 0.85 1.06 (P) 1.00 0.93 1.28 1.00 1.12 1.12
S15-143 114 (4.5) 1.43 776 (174.4) P 0.97 1.47 (P) 1.22 0.98 1.22 0.99 1.17 1.17
Minimum 0.49 0.68 1.06 0.89 0.98 0.97
Maximum 1.32 1.16 2.42 1.21 1.29 1.23
Average 0.90 0.95 1.52 1.01 1.11 1.08
Coefficient of variation 0.30 0.15 0.29 0.084 0.082 0.083
*
Vexp is total failure load, including weight of specimen outside distance d from column face and weight of test equipment on top of specimen.

Observed failure mode: F is flexural failure; P is punching failure.

ρfs is proposed limiting reinforcement ratio, calculated using Eq. (12a), ρfs is taken to be 0.007 for proposed Simplified Method as shown by Eq. (12b).
§
For ρ/ρfs > 1.0, the predicted failure mode is punching; for ρ/ρfs ≤ 1.0, the predicted failure mode is flexure. ρfs in Column (7) is for proposed Standard Method and calculated using
Eq. (12a).

way to failure. Crack widths were measured by using a crack


detector until near failure.

TEST RESULTS AND DISCUSSIONS


Failure loads and crack patterns
The failure loads Vexp of the 12 specimens are summarized
in Table 2. Each failure load includes the self-weight of the
slab outside the perimeter measured at d away from a column
face and the weight of test equipment placed on top of the
slab. Those slabs that failed abruptly with a sudden drop in
their load-deflection curves are indicated to have failed in
punching mode (P). Those slabs that failed in a more ductile
manner or with widespread yielding of reinforcement prior
to punching failure are indicated to have failed in flexural
(F) mode. All the slabs finally failed in what appears to be
punching failure, even though the slabs might have failed
earlier in flexural mode.
Figure 3 shows the photographs of the crack patterns
at the failure of the S11-series. The crack patterns of the Fig. 3—Crack patterns at failure: (a) S11-028; (b) S11-050;
other eight slabs are shown in Fig. 4. Figure 3 shows fairly (c) S11-090; and (d) S11-139.
clearly that slabs with low reinforcement ratios would have
different crack patterns at failure compared to slabs with the final circumferential crack that comes to the surface from
high reinforcement ratios. Typical crack patterns of slabs the internal inclined shear cracks.
failing in normal punching can be seen in Fig. 3(c) and (d), If the reinforcement ratio is low, however, the circular-
which show Specimens S11-090 and S11-139, respectively. fan-type of crack pattern may not form. Figure 3(a) shows
The crack pattern in each of these cases comprises a set of the crack pattern of Slab S11-028; it can be seen that the
closely spaced radial cracks or circular-fan-type cracks with crack pattern forms straight-line cracks nearly parallel

142 ACI Structural Journal/January 2018


Fig. 4—Crack patterns of S13 and S15 series at failure.
to column lines in both directions with perhaps one diag-
onal crack from the column corner toward a corner of the
slab. The parallel line crack pattern shows that widespread
yielding of the reinforcement has occurred. In the end, the
final failure of the slab would still be a punching failure with
the final circumferential crack occurring very close to the
column. The failure load is at a load no greater than the load
that caused earlier yielding of the flexural reinforcement.
Similar behavior and crack patterns can also be seen in other
specimens with low reinforcement ratios such as S13-028
(ρ = 0.28%) and S15-028 (ρ = 0.28%), as shown in Fig. 4.

Deflections
The load-deflection curves of the 12 specimens are shown
in Fig. 5. Each curve shows the average of four deflection
points located 100 mm (3.9 in.) from the slab edges. Before
cracking, the relationship between the load and deflection is
linear. After the first circumferential crack formed, the slope
of the load-deflection curve would change slightly. Upon
further loading, the change of slope becomes increasingly
more significant as the flexural stiffness of the slab drops
further due to more cracking or widening of cracks. As
expected, the flexural stiffness of the slab with a higher rein-
forcement ratio will degrade less after cracking—that is, its
load-deflection curves have steeper slopes compared to slabs
with low reinforcement ratios.
Figure 5 also shows that slabs with lower reinforcement
ratios are more ductile than slabs with higher reinforcement
ratios. Upon reaching the maximum load, any further load
increment to Specimen S11-028, S13-028, and S15-028
(very low reinforcement ratios) produces no additional Fig. 5—Load-deflection curves of 12 slabs: (a) S11-series;
increase in resistance, but their deformations continue to (b) S13-series; (c) S15-series.
increase until the final failure at the end. The slabs with
the highest reinforcement ratio (S11-139, S13-143, and Strains in flexural reinforcements
S15-143) are the most brittle. Nevertheless, the slabs with Figure 6 shows the summary of steel strain distributions
higher reinforcement ratios also have higher failure loads. at the ultimate load stage in the S11-series, S13-series, and
However, the influence of reinforcement ratio on punching S15-series. From Fig. 6, it can be seen that in slabs with
shear strength is neglected in the ACI 318-14.1 higher reinforcement ratios (ρ ≥ 0.9% or Sxy-090 and

ACI Structural Journal/January 2018 143


Fig. 6—Steel strains in several reinforcing bars near columns at failure loads for: (a) S11 series slabs; (b) S13 series slabs;
and (c) S15 series slabs.
Sxy-143), most of the steel strains drop considerably at by assuming a square column of equivalent area. The influ-
locations beyond 1.5h away from the column face. In slabs ence of flexural reinforcement and size effect are neglected
with lower reinforcement ratios (ρ ≤ 0.5% or Sxy-028 and in the ACI Code equations. ACI 318-141 does not limit the
Sxy-050), the initial failure mode is flexure, and most of the maximum concrete strength fc′ but the value of √fc′ used to
steel strains can remain high or greater than the yield strain calculate shear strength is limited to 100 psi (8.3 MPa).
even at locations near the edge of the slabs. Thus, in slabs
that fail in pure punching, the steel strains reach the yield Eurocode 22
strain only near the column. According to the Eurocode 2,2 the punching shear resis-
The consequence of the aforementioned behavior is that tance VRd,c of a slab without shear reinforcement is given in
the reinforcement that should be considered effective for Eq. (4)
resisting punching load are those within approximately 1.5h
from the column faces or within a width of c2 + 3h for an VRd,c = 0.18k(100ρfc′)1/3u1d (SI units)
interior column. (4)
VRd,c = 5k(100ρfc′)1/3u1d (U.S. units)
DESIGN CODES AND EXISTING DESIGN METHODS
In this section and the next section, the ACI 318, Euro- where k is the size effect coefficient, = 1 + 200 / d ≤ 2.0 (SI
code 2 methods, and some existing design methods proposed units) or = 1 + 8 / d ≤ 2.0 (U.S. units). The ρ is the average
by researchers are discussed briefly.
flexural reinforcement ratio ( ρ = ρx ρ y ≤ 0.02 ). The crit-
ical shear perimeter u1 is located at a distance 2d away
ACI 318-141
from the column faces and it has round corners. Eurocode
According to ACI 318-14,1 the punching shear strength Vc
2 neglects the effect of column rectangularity in symmetri-
of slabs without shear reinforcement can be determined from
cally loaded slabs.
the lowest of the following expressions
Existing design methods
1
Vc = (2 + 4 / β) f c′bo d (SI units) Teng et al.3 investigated the effects of slab openings
12 (1) and column rectangularity in normal-strength reinforced
Vc = (2 + 4 /β) f c′bo d U.S. units)
(U concrete slabs. They proposed a design equation for the
punching shear strength Vc, as shown in Eq. (5)
1
Vc = (α s d /bo + 2) f c′bo d (SI units) Vc = 0.6kCR(100ρfc′)1/3bod (SI units)
12 (2) (5)
Vc = (α s d /bo + 2) f c′bo d (U.S. units) Vc = 16.6kCR(100ρfc′)1/3bod (U.S. units)

where
1
Vc = f c′bo d (SI units)
3 (3) kCR = (b2/b1)1/4 < 1.0 (6)
Vc = 4 f c′bo d (U.S. units)
where b1 and b2 are the longer and shorter sides of the crit-
where β is the ratio of long to short sides of the column; αs ical shear perimeter, respectively. The use of b1 and b2 in the
is taken to be 40, 30, and 20 for interior, edge, and corner column rectangularity factor, kCR, rather than c1 and c2, is
columns, respectively; and bo is the length of the critical to take into consideration the thickness of the slab in influ-
perimeter located at 0.5d away from the column face. For encing the distribution of stresses around the column. The
a circular column, the critical section can also be defined critical shear perimeter bo has square corners for both square
and circular columns and is located at a distance d/2 away

144 ACI Structural Journal/January 2018


from column faces. For a circular column, an equivalent the slab to fail in flexural mode. For slabs with ρ < ρfs, Peiris
square column of the same area can be used for bo calcu- and Ghali5 assume that Vflex to be a linear function of ρ. So,
lations. The equation was verified3 with slab data having a the failure load for slabs with ρ < ρfs is then equal to (ρ/ρfs)Vc.
broad range of relevant parameters. However, the applica- The key lies in the term ρfs. The Vflex can be calculated by
bility of the equation for slabs with very low reinforcement using the yield line theory. For some simple cases, Vflex can
ratios and very thick slabs has not been verified thoroughly. be approximated5 by 8m, where m is moment capacity per
Muttoni6 introduced the Critical Shear Crack Theory unit width (ACI 318). By setting Vflex = Vc(ACI), the ρ becomes
(CSCT) for calculating the punching shear strength of slabs ρfs, and the ρfs can be obtained. So
without shear reinforcement. The method assumes that the
punching shear strength VRd of a slab depends on the rotation 8m = 4 f c′bo d , where m = 0.95 fyd2ρ (9)
ψ of the slab at failure, as well as crack width and aggre-
gate interlock along the critical shear crack. The CSCT Then, solving for ρ or ρfs
also considers that the rotation ψ of the slab depends on the
applied punching shear force V and a few other parameters. 4 f c′bo d
Failure is reached when the applied punching shear force V ρ fs = (U.S. units) (10)
is equal to the punching shear strength of the slab VRd. 8 × 0.95 f y d 2
The method requires an iterative solution to two equations:
1) the relationship between the punching shear strength VRd DERIVATION OF PROPOSED EQUATION
and rotation ψ; and 2) the relationship between the rotation ψ The authors’ proposed equation shown in Eq. (15) is an
and applied shear force V. Equations (7) and (8) show these extension of the punching shear strength formula that was
two equations previously introduced by Teng et al.,3 shown earlier in
Eq. (5). New reduction factors to consider low reinforce-
0.75 ment ratio and size effect are added.
VRd = f c′bo d (SI units)
ψd
1 + 15 Factor for low reinforcement ratio, kRR
d go + d g
(7) The effect of low reinforcement ratio is treated in the same
9 way as in Peiris and Ghali’s approach. If the reinforcement
VRd = f c′bo d (U.S. units)
ψd ratio is less than ρfs, then the failure mode changes from
1 + 15
d go + d g punching to flexural mode and a reduction factor for low ρ
is activated. A suitable yield line pattern is chosen based on
1.5 the authors’ experimental results to determine the flexural
r fy  V  strength Vflex of slabs under punching load. Then, by setting
ψ = 1.5 s (8)
d Es  V flex  the Vflex to be equal to Vc of Eq. (5) and taking into consid-
eration the reduction factors for size effect and column rect-
where rs is the radius of the slab (equal to half of a slab angularity, the reduction factor kRR for the effect of low rein-
length or 0.22L, where L is column-to-column span). Vflex is forcement ratio can be written as (the complete derivation of
the shear force that will cause flexural failure of the slab. bo the factor kRR is given in the Appendix A,* which is included
is calculated at a distance d/2 away from column edges with in the electronic version of the paper)
round corners. dgo is set to be 16 mm and dg is the maximum
aggregate size. Muttoni6 states that the size effect in CSCT is kRR = (ρ/ρfs)1/6 (11)
a function of rs or slab span rather than the effective depth d.
The CSCT method can become complex, as it requires where ρ is the provided reinforcement ratio; and ρfs is given by
an iterative procedure to solve for VRd. In addition, every
slab geometry, reinforcement ratio, and loading condition 3/ 2
 0.6kCR k SZ f c′1/ 3bo d 
would lead to different yield line pattern, requiring yield ρ fs = (0.01)  2
(100)  (SI units)
line analysis to obtain Vflex. Consequently, for general design  α o (0.95 f y d ) 
purposes, Vflex can be approximated4,6 and taken equal to 8m, 3/ 2
(12a)
16.6kCR k SZ f c′1/ 3bo d 
where m is the moment capacity per unit width. ρ fs = (0.01)  (100)  (U.S. units)
2
Peiris and Ghali5 proposed a straightforward method  α o (0.95 f y d ) 
to improve the ACI Code method by introducing a reduc-
tion factor, ρ/ρfs, to make it safer for slabs with low flexural where kCR and kSZ are the reduction factors for column rect-
reinforcement ratios. ρfs is the limiting reinforcement ratio. angularity ratio and size effect, respectively (they will be
If the provided reinforcement ρ is reduced to below ρfs, the discussed in the following). αo = [2(c1 + c2)/r + 2π] with r
slab is assumed to fail in flexural mode instead of the usual being the distance from the column face to the point load and
punching mode. For slabs with provided reinforcement ρ > it can be taken as 0.2 times span length.
ρfs, the punching shear strength of the slab Vc is determined
by the usual ACI equations (Eq. (1) to (3)). If the provided ρ *
The Appendix is available at www.concrete.org/publications in PDF format,
in the slab is reduced below ρfs, the punching shear strength appended to the online version of the published paper. It is also available in hard copy
Vc becomes equal to Vflex—that is, the shear force that causes from ACI headquarters for a fee equal to the cost of reproduction plus handling at the
time of the request.

ACI Structural Journal/January 2018 145


(12 in.), kSZ = 1.0. So, for the majority of slabs in practical
structural design, the size effect factor can be disregarded.
For the column rectangularity ratio, β (= c1/c2), greater
than 1.0, the corresponding reduction factor kCR in Eq. (6),
which was empirically derived, is retained with the exponent
adjusted to 1/3 to suit more data

kCR = (b2/b1)1/3 (14)

Figure 7 shows the graph depicting the normalized shear


stress at failure Vexp/{bod(100ρfc′)1/3} against β (the ratio of
long to short sides of the column, or c1/c2) for the authors’
current slab specimens and other existing slab data from the
literature.3,8-11 Each of selected set of data comprises slab
specimens with different column aspect ratios. The authors’
slab specimens were made of high-strength concrete, while
those from the literature were made of normal-strength
concrete. It can be seen from the trend line for each set of
data that the normalized shear stress decreases as the column
aspect ratio increases. The trend lines are all similar to each
other, including the one for the authors’ current high-strength
concrete slabs.
Fig. 7—Performance of proposed column rectangularity Figure 7 essentially shows the performance of the column
factor kCR. rectangularity factor, kCR, in comparison with experimental
data. It can be seen that kCR represents the trend line of the
Note, however, that the general formula for ρfs given by experimental data (including high-strength concrete slabs)
Eq. (12a) can also be simplified considerably. By assuming very well. The way ACI treats rectangular column is by the
typical values encountered in practice (refer to Appendix A), use of the β coefficient as shown in Eq. (1). It cannot be
the simplified value of ρfs can be taken as shown together in Fig. 7 because ACI uses √fc′ instead of
fc′1/3 in its equation. Nevertheless, the ACI method of treating
ρfs = 0.007 or 0.7% (12b) rectangular columns as shown in Eq. (1) also represents the
experimental data very closely.
Thus, the authors now have two proposed methods: 1) the
general method, which uses the ρfs of Eq. (12a); and 2) the Proposed equation
simplified method, which uses ρfs of Eq. (12b). The simplified The general form of the proposed equation can be written as
method will give equally accurate predictions of punching
failure loads but less accurate predictions of failure modes. VC = 0.6kRRkCRkSZ(100ρfc′)1/3bod (SI units)
(15)
Size effect and column rectangularity ratio VC = 16.6kRRkCRkSZ(100ρfc′)1/3bod (U.S. units)
Various approaches have been employed to study the size
effect in concrete structures. A fracture mechanics-based where kRR, kCR, and kSZ are the reductions factors for rein-
approach pioneered by Bažant et al.7 leads to a reduction forcement ratio, column rectangularity ratio, and size effect
factor that is proportional to d–1/2. However, due to the (refer to Eq. (11), (13), and (14)). ρ is As/bd and should not be
variability of test results, properties of concrete members, taken greater than 2.5%. Note that in most cases, the reduc-
and the influence of reinforcement, an empirical approach tion factors need not be calculated, as they will be equal to 1.
is often used, and this can lead to a reduction factor that is For commonly used slabs supported on square columns
proportional to d–1/3 to d–1/4 or others. (c1/c2 =1) with reinforcement ratios ρ greater than 0.7%, and
By observing the available experimental data and prac- effective depths d of 300 mm (12 in.) or less, the proposed
tical design considerations, the size-effect reduction factor equation in Eq. (15) becomes simple as follows
kSZ in the following is adopted for the proposed equation
VC = 0.6(100ρfc′)1/3bod (SI units)
kSZ = (300/d)1/2 ≤ 1.0 (SI units) (16)
(13) VC = 16.6(100ρfc′)1/3bod (U.S. units)
kSZ = (12/d)1/2 ≤ 1.0 (U.S. units)
COMPARISON WITH EXPERIMENTAL RESULTS
kSZ is chosen to be applicable for an effective depth d > Comparison with authors’ 12 new high-strength
300 mm (12 in.) simply because the current basic equation concrete slab specimens
for Vc is safe for d of up to 300 mm (12 in.). For d ≤ 300 mm For the purpose of these comparisons, all the load factors,
materials safety factors, and strength reduction factors are

146 ACI Structural Journal/January 2018


all set equal to 1.0. Shown in Table 2 are the punching shear able method is expected to perform well in all the different
strengths of the 12 high-strength concrete slabs presented ranges or divisions.
earlier in this paper and the predictions of each of the methods Influence of reinforcement ratio—Figures 8(a) and (b)
discussed previously. It can be seen in Column (9) that the show the comparison between the experimental failure load
predictions of the ACI Code method are unconservative Vexp and ACI 318-141 and Eurocode 22 predictions (Vcalc),
(Vexp/Vcalc < 1.0) for slabs with low reinforcement ratios of respectively. The ratio of the experimental failure load to the
0.28% or 0.5%, as expected. For slabs with higher ρ, such as calculated punching shear strength (Vexp/Vcalc) is plotted in
Sxy-090 and Sxy-143 series, the ACI predictions are conser- the y-axis, with the x-axis being three design parameters:
vative. The corrections to ACI method that was introduced ρ, d, and fc′. Separate statistical analysis is done for each
by Peiris and Ghali5 (Column (11)) make the corrected ACI of the divisions A1, A, B, and C. The horizontal dashed
method very conservative for low reinforcement ratio cases. line in each division represents the average values of the
For the Eurocode 2 predictions shown in Column (10), the Vexp/Vcalc for that division. Figures 9(a), (b), and (c) show the
predictions are reasonably conservative, but it cannot take comparisons of the experimental results with the predictions
into account the reduction in punching shear strengths of of CSCT,6 Peiris-Ghali,5 and the authors’ general methods,
slabs supported on rectangular columns (S15 series espe- respectively. The statistical analysis results for each division
cially) and of slabs with very low reinforcement ratio of are summarized in a tabulated format in Table 3.
< 0.3%, such as S11-028, S13-028, and S15-028. Muttoni The performance of the ACI equations with respect
et al.’s6 CSCT method (Column 12) performs very well for to reinforcement ratio is shown in the leftmost chart in
these 12 slabs, with a coefficient of variation of only 8.4% Fig. 8(a). It can be seen that as the reinforcement ratio
and an average of Vexp/Vcalc of 1.01. increases, the average of Vexp/Vcalc increases as well. Simi-
The authors’ proposed general prediction method (Column larly, Table 3 shows for the ACI318-14 predictions that as
13) is the most accurate, with an average of Vexp/Vc of 1.09 the reinforcement ratio increases from Group A1 (Column
and a coefficient of variation of only 8.2%. The authors’ (4) or ρ < 0.6%) to C (Column (7) or ρ > 2%), the average
proposed general prediction method can also predict the of the Vexp/Vcalc increase from 0.86 to 1.61. This is caused by
failure modes of the slabs accurately. Column (8) shows the the neglect of the reinforcement ratio in the Vc formulas in
values of ρ/ρfs from the standard ρfs equation (Eq. 12(a)). If the ACI 318-14. Note also that for the region A1(ρ <0.6%)
the ρ/ρfs is less than 1.0, the corresponding slab would fail in both Fig. 8(a) and Table 3, the average is 0.86 and most
in flexural mode (F); otherwise, if ρ/ρfs is equal or greater of the points are below the 1.0 line. This confirms the earlier
than 1.0, the slab is expected to fail in punching (P). Column conclusion from the 12 HSC slabs that the ACI 318 method
(6) shows the actual failure modes of the slabs. Comparing for punching shear is unconservative for slabs with low rein-
Column (8) and Column (6), it can be seen that the authors’ forcement ratio, specifically for ρ less than approximately
general method predicts the actual failure modes very accu- 0.6% or 0.7%.
rately—that is, all the failure modes are predicted correctly. The correction procedure introduced by Peiris-Ghali to take
In fact, both the general and the simplified methods predict care of possible flexural failure prior to punching failure is
the failure modes of the 12 HSC slabs correctly. effective for the intended A1 region. From Table 3, in Peiris-
Ghali rows, the average of Vexp/Vcalc in division A1 increases
Comparison with slab database of 367 specimens from 0.86 (ACI average) to a healthy 1.15. However, the
Table 3 summarizes the punching shear-strength predic- maximum Vexp/Vcalc increases from 1.19 to 2.42. This shows
tions of the methods mentioned previously for the 367 that the correction can also lead to overconservatism of the
slabs in the database. The material properties and complete method. Both the Eurocode 2 (EC2) method and Muttoni
details of the 367 slab data and the detailed comparisons of et al.’s CSCT method give good predictions with the CSCT
the failure load of each slab with the predictions of all the having slightly lower coefficients of variations (average
methods are given in Table A1 in Appendix B. COV, that is; average of the four divisions = 0.16 compared
The current database of 367 slab data was from the authors’ to 0.18 for EC2). Figure 9(a) shows that CSCT predictions
earlier database3 with the addition of slab data from the data- in terms of reinforcement ratio has narrower band (better)
base of the ACI Subcommittee 445-C12 and new slab data compared to Fig. 8(b) for EC2. Note, however, that both
from the literature shown in Table A1 in Appendix B. All methods have a minimum Vexp/Vcalc of 0.64 for EC2 and 0.57
the slab data in the current database is selected to make sure for the CSCT method. These are quite low.
that they satisfy all the relevant ACI design requirements. The authors’ proposed method can be considered the best
Slabs having unusual geometries, loading arrangements, with respect to reinforcement ratio. The combined average
or boundary conditions were excluded, as their punching of Vexp/Vcalc of the four divisions (sum of four averages
strengths would be quite different from those of normal slabs divided by 4) is approximately 1.2, with a combined average
in typical construction and they deserve special treatment of COV of approximately 0.15. The leftmost chart of Fig. 9(c)
their own. shows the authors’ predictions using the general method.
The data are grouped according to three parameters (rein- The average of one division is close to the averages of the
forcement ratio, effective depth, and compressive strength) other three divisions, and the data are in a reasonably narrow
and divided into three or four different divisions or ranges as band, indicating a low COV.
detailed in Fig. 8 and 9 and also in Table 3. A good and reli- Influence of effective depth—The middle charts in Fig. 8
and 9 and Columns 8 to 10 of Table 3 show the statistical

ACI Structural Journal/January 2018 147


Table 3—Comparison of experimental and calculated failure load by all methods
Vexp/Vcalc Reinforcement ratio ρ, % Effective depth d, mm Concrete strength fc′, MPa
[A1] [A] [B] [C] [A] [B] [C] [A] [B] [C]
ρ< 0<ρ≤ 1.0 < ρ ≤ ρ> d ≤ 100 100 < d < d ≥ 300 fc΄ ≤ 50 50 < fc΄ ≤ fc΄ > 90
Division, units All data 0.6% 1.0% 2.0% 2.0% mm 300 mm mm MPa 90 MPa MPa
No. of data: 367 57 160 175 41 114 247 6 279 63 25
(1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) (12) (13)
Minimum 0.49 0.49 0.49 0.76 1.06 0.70 0.49 0.64 0.51 0.59 0.49
Maximum 2.32 1.19 1.78 1.92 2.32 2.32 1.98 1.10 2.32 1.79 1.98
Vexp/Vcalc
ACI 318-14 Average 1.26 0.86 1.08 1.37 1.61 1.38 1.22 0.86 1.28 1.22 1.17
Coefficient
0.249 0.233 0.258 0.171 0.167 0.223 0.247 0.193 0.240 0.236 0.366
of variation
Minimum 0.64 0.68 0.68 0.64 0.83 0.79 0.64 0.79 0.69 0.64 0.68
Maximum 2.04 1.32 1.77 2.01 2.04 2.04 1.48 0.99 2.04 1.48 1.23
Vexp/Vcalc
Eurocode 2 Average 1.14 1.02 1.11 1.17 1.21 1.26 1.09 0.90 1.16 1.10 1.02
Coefficient
0.174 0.169 0.175 0.183 0.179 0.158 0.156 0.085 0.171 0.181 0.138
of variation
Minimum 0.57 0.64 0.64 0.57 0.76 0.65 0.57 0.68 0.64 0.57 0.89
Maximum 1.82 1.37 1.37 2.19 1.38 1.82 1.48 1.21 1.82 1.31 1.23
Vexp/Vcalc
CSCT Average 1.04 1.02 1.03 1.06 1.02 1.07 1.02 0.86 1.04 1.02 1.04
Coefficient
0.152 0.17 0.145 0.198 0.129 0.156 0.145 0.22 0.151 0.174 0.094
of variation
Minimum 0.75 0.76 0.75 0.76 1.06 0.90 0.76 0.75 0.75 0.76 0.94
Maximum 2.42 2.42 2.42 2.31 2.32 2.32 2.42 1.10 2.32 1.79 2.42
Vexp/Vcalc
Peiris-Ghali Average 1.31 1.15 1.19 1.37 1.61 1.42 1.28 0.89 1.32 1.23 1.47
Coefficient
0.214 0.27 0.217 0.177 0.167 0.185 0.215 0.152 0.201 0.217 0.282
of variation
Minimum 0.73 0.73 0.73 0.76 0.83 0.86 0.73 1.16 0.73 0.83 0.97
Vexp/Vcalc Maximum 1.69 1.55 1.67 1.69 1.50 1.69 1.67 1.28 1.69 1.56 1.43
proposed
Simplified Average 1.21 1.18 1.22 1.22 1.16 1.28 1.18 1.23 1.21 1.24 1.16
Method Coefficient
0.151 0.151 0.151 0.151 0.153 0.127 0.158 0.031 0.150 0.160 0.131
of variation
Minimum 0.76 0.81 0.81 0.76 0.83 0.86 0.76 1.13 0.76 0.83 0.98

Vexp/Vcalc Maximum 1.69 1.49 1.67 1.69 1.50 1.69 1.67 1.24 1.69 1.56 1.43
proposed Average 1.20 1.13 1.20 1.22 1.16 1.27 1.17 1.19 1.20 1.23 1.17
General Method
Coefficient
0.152 0.143 0.151 0.153 0.153 0.129 0.158 0.035 0.151 0.164 0.125
of variation

Notes: Regions A1, A, B, C correspond to the regions in Fig. 8 and 9. 1 mm = 0.0394 in., 1 MPa = 145.0 psi.

analysis results as they are influenced by the effective Both the authors’ simplified and general methods show
depth, d. It can be seen from the middle chart in Fig. 8(a) very good agreements with the experimental results. The
that the average of Vexp/Vcalc of the ACI 318-141 becomes average lines from the authors’ general method across the
lower as the effective depth d increases. The ACI method different divisions of effective depths as shown in Fig. 9(c)
can be unconservative for slabs having an effective depth of are very close to each other, indicating a very good match
greater than 300 mm (12 in.). Obviously, this is caused by between the equations and actual failure behavior of the
the neglect of size effect in the ACI equations. The Peiris- slabs. The average of the three divisions is approximately
Ghali method, which is a correction to ACI method, does not 1.2 and the average COV is approximately 0.11 for both the
address size effect. simplified and standard equations.
The EC22 equation, which has a size effect term, also does Influence of compressive strength—The third charts (right-
not perform well when d is greater than 300 mm (12 in.) most) in Fig. 8(a) shows the performance of ACI method
(refer to the middle chart in Fig. 8(b)). The CSCT method in terms of compressive strength fc′. It can be seen that the
also does not predict well the punching shear strengths of average of Vexp/Vcalc in one division or range (represented
slabs that are thicker than 300 mm (12 in.). by the horizontal dashed line) is nearly the same as that of

148 ACI Structural Journal/January 2018


Fig. 8—Failure load predictions of 367 test data by: (a) ACI 318-14; and (b) Eurocode 2 2004.

Fig. 9—Failure load predictions of 367 test data by: (a) CSCT; (b) Peiris and Ghali; and (c) Authors’ proposed General Method.
the other divisions. This is true for the other four methods general and simplified methods with the COVs of 0.125 and
shown in Fig. 8(b) and 9(a), (b), and (c). However, the ACI 0.131, respectively, followed by the EC2, Peiris-Ghali, and
method has the widest spread of results with combined COV ACI methods.
across the three divisions of 0.28. The CSCT method provides Other parameters, failure modes, and summary—It is
the best predictions with a combined average of Vexp/Vcalc of understood that influencing parameters can be interrelated
1.03 and with a combined COV of 0.14. The authors’ simpli- and they may also influence each other. These are consid-
fied and general prediction equations both have a combined ered as much as possible. Other parameters that might influ-
average of 1.2 and a combined COV of 0.15. Both CSCT ence the punching shear strength include column size to slab
and the authors’ methods can be considered equally accu- thickness effect (c/d or bo/d), maximum aggregate size da,
rate. Table 3 (Column 13, for fc′ > 90  MPa [13,000 psi]) and also shear span-depth ratio (r/d or 0.2L/d). However,
shows that the CSCT method predicts the punching shear those effects were insignificant.
strength of HSC slabs (fc′ > 90 MPa [13,000  psi]) very From the aforementioned comparative study, it can be
well with an average of 1.04 and a COV of 0.094. The next seen that the authors’ proposed General Method and the
best method for fc′ > 90 MPa [13,000 psi] is the authors’ Simplified Method are accurate and reliable. The General

ACI Structural Journal/January 2018 149


Method predicts failure modes more accurately than the AUTHOR BIOS
Simplified Method (refer to Table A1 in Appendix B). Of the ACI member Susanto Teng is an Associate Professor at the School of Civil
and Environmental Engineering, Nanyang Technological University (NTU),
27 slabs that failed in flexure, the general method correctly Singapore. He is a member of ACI Committee 435, Deflection of Concrete
predicted 20 of them (74% correct) while the simplified Building Structures, and Joint ACI-ASCE Committees 421, Design of Rein-
method predicts 21 correctly (78%). Of the 340 slabs failing forced Concrete Slabs, and 445, Shear and Torsion. His research interests
include shear in beams, slab-column connections, finite element modeling,
in punching, the general method predicts mode of failure and tall buildings.
correctly for 316 (93%) of them compared to 274 (81%) for
the simplified method. Nevertheless, both methods predict Khatthanam Chanthabouala is a PhD Candidate in the School of Civil
and Environmental Engineering at NTU. His research interests include
the punching shear strength equally accurately and they punching shear strength of high-strength concrete and steel fiber-reinforced
give the closest predictions to the experimental failure loads concrete slabs.
compared to other methods. The simplified method is espe-
Darren T. Y. Lim is a PhD Candidate in the School of Civil and Environ-
cially very simple to use. mental Engineering at NTU. His research interests include concrete tech-
nology and shear behavior of large concrete beams.
CONCLUSIONS
Rhahmadatul Hidayat is a Civil Engineer with Mott MacDonald Consul-
Based on the authors’ experiments and the accompanying tants, Singapore. He received his master’s degree from NTU. His research
discussion as well as the comparative study of six different interests include tunnel construction for the underground Mass Rapid
methods against the database of 367 slabs, the following Transit system in Singapore.
conclusions can be made; they are applicable to normal-
ACKNOWLEDGMENTS
strength as well as high-strength concrete slabs: This research was made possible by the funding provided by the National
1. Under a punching load, a slab provided with a low rein- Research Foundation (NRF) of Singapore and the support of Nanyang
forcement ratio (approximately less than 0.6 to 0.7%) may Technological University, Singapore.
fail in flexure first before it fails in what looks like punching
failure in the end. The failure mode will be ductile and REFERENCES
1. ACI Committee 318, “Building Code Requirements for Structural
the failure load will be the load that causes yielding of the Concrete (ACI 318-14) and Commentary (ACI 318R-14),” American
majority of the reinforcement. The failure load will be lower Concrete Institute, Farmington Hills, MI, 2014, 519 pp.
than predicted by ACI 318-14 equations for punching shear. 2. CEN, “Eurocode 2—Design of Concrete Structures: Part 1-1—
General Rules and Rules for Buildings,” EN 1992-1-1, Brussels, Belgium,
2. Slabs failing in pure punching shear mode will tend to 2004, 225 pp.
form circular-fan-type crack patterns at failure, while those 3. Teng, S.; Cheong, H. K.; Kuang, K. L.; and Geng, J. Z., “Punching
failing in flexural mode will tend to form cracks parallel to Shear Strength of Slabs with Openings and Supported on Rectangular
Columns,” ACI Structural Journal, V. 101, No. 5, Sept.-Oct. 2004,
column lines. In this case, the final circumferential crack pp. 678-687.
tends to be very close to the column. 4. Guandalini, S.; Burdet, O. L.; and Muttoni, A., “Punching of Slabs
3. The higher the amount of reinforcement provided, the with Low Reinforcement Ratios,” ACI Structural Journal, V. 106, No. 1,
Jan.-Feb. 2009, pp. 87-95.
higher the failure load. However, only those reinforcements 5. Peiris, C., and Ghali, A., “Flexural Reinforcement Essential for
within the width of 1.5h from a column face (or total width Punching Shear Resistance of Slabs,” Recent Development in Reinforced
of c2 + 3h for an interior column) will be fully effective Concrete Slab Analysis, Design, and Serviceability, SP-287, M. Mahamid
and F. Malhas, eds., American Concrete Institute, Farmington Hills, MI,
in resisting punching shear stresses and, therefore, can be 2012, pp. 1-16.
considered in influencing the punching shear strength. 6. Muttoni, A., “Punching Shear Strength of Reinforced Concrete Slabs
4. The ACI 318 method is unconservative for slabs with without Transverse Reinforcement,” ACI Structural Journal, V. 105, No. 4,
July-Aug. 2008, pp. 440-450.
low reinforcement ratios (especially for ρ < 0.7%). The 7. Bažant, Z. P.; Yu, Q.; Gerstle, W.; Hanson, J.; and Ju, J. W., “Justifi-
correction procedure proposed by Peiris and Ghali fixes cation of ACI 446 Proposal for Updating ACI Code Provisions for Shear
the problem but it may lead to over-conservatism. The Design of Reinforced Concrete Beams,” ACI Structural Journal, V. 104,
No. 5, Sept.-Oct. 2007, pp. 601-610.
EC2 method can be unconservative for slabs with very low 8. Rosenthal, I., “Experimental Investigation of Flat Plate Floors,” ACI
reinforcement ratio and supported on rectangular columns. Journal Proceedings, V. 56, No. 12, Dec. 1959, pp. 153-166.
5. The ACI 318, EC2, and CSCT methods may not be 9. Moe, J., “Shearing Strength of Reinforced Concrete Slabs and Foot-
ings under Concentrated Loads,” Bulletin D47, Portland Cement Associa-
safe for slabs with an effective depth of more than 300 mm tion, Research and Development Laboratories, Skokie, IL, 1961, 130 pp.
(12 in.), even though EC2 and CSCT do consider size effect. 10. Hawkins, N. M.; Fallsen, H. B.; and Hinojosa, R. C., “Influence
ACI needs to consider the inclusion of size effect into the of Column Rectangularity on the Behaviour of Flat Plate Structures,”
Cracking, Deflection, and Ultimate Load of Concrete Slab Systems, SP-30,
equation. American Concrete Institute, Farmington Hills, MI, 1971, pp. 677-720.
6. The authors’ general method and simplified method 11. Oliveira, D. R. C.; Regan, P. E.; and Melo, G. S. S. A., “Punching
both have been shown to be most accurate and reliable with Resistance of RC Slabs with Rectangular Columns,” Magazine of Concrete
Research, V. 56, No. 3, 2004, pp. 123-138. doi: 10.1680/macr.2004.56.3.123
uniform averages of Vexp/Vcalc and COV across different 12. Ospina, C. E.; Birkle, G.; Widianto; and Kuchma, D., “ACI
ranges of design parameters. The simplified method is espe- 445C - Punching Shear - Collected Databank,” 2011, https://nees.org/
cially simple to use. resources/3660/. (last accessed Dec. 18, 2017)

150 ACI Structural Journal/January 2018


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 115-S12

Biaxial Interaction and Load Contour Method of Reinforced


Concrete T-Shaped Structural Walls
by Tae-Sung Eom, Hye-Sung Nam, and Su-Min Kang

Reinforced concrete structural walls with T-shaped cross sections Alternatively, the equivalent eccentricity method, which
(or T-shaped walls) have been used as an efficient lateral was originally developed for biaxially loaded columns, can
force-resisting system for building structures. Such T-shaped walls be used for the biaxial design of nonplanar walls (Wight and
are subjected to axial compression and combined bending moments MacGregor 2012; Furlong et al. 2004; MacGregor 1973).
about two orthogonal axes. In the present study, a straightforward
In this method, a fraction between 0.4 and 0.8 times the
design method for biaxially loaded T-shaped walls is developed
weak-axis moment is added to the strong-axis moment, and
by modifying the existing load contour method. First, a strain
compatibility section analysis method that can estimate the biaxial the wall is then designed for the axial load and increased
bending resistances of arbitrary wall sections is developed and its strong-axis moment treated as a case of uniaxial bending
validity is verified through comparisons with test results. Then, and compression. More straightforward and approximate
a parametric study is performed to investigate the interaction of methods treating the biaxial bending and compression
biaxial moments at a constant axial load in T-shaped walls. The directly are the load contour method and reciprocal load
parametric results show that, due to the unsymmetrical geometry method (Bresler 1960; Parme et al. 1966; Portland Cement
of T-shaped sections, the biaxial interaction depends significantly Association 1966, 2013; Furlong et al. 2004; Cedolin et al.
on the direction of moments and the magnitude of axial compres- 2008; El Fattah et al. 2013). Particularly in the load contour
sion. Based on the results, the non-dimensional contour equations method, the three-dimensional (3-D) interaction surface of
of biaxial moments for T-shaped walls are proposed and a design
biaxial moments and axial load is converted into a family of
procedure of T-shaped walls is established.
two-dimensional (2-D) load contours representing the rela-
Keywords: biaxial interaction; biaxially loaded walls; load contour tionships between the moments about two axes at a constant
method; reinforced concrete; structural wall; T-shaped walls. axial load. Thus, the load contour method is convenient to
understand the biaxial interaction more intuitively. In addi-
INTRODUCTION tion, through the process of normalization, the load contour
In high-rise buildings, nonplanar walls with various method, which was originally developed for biaxially loaded
section geometries, such as T-shaped, C-shaped, and columns with solid sections such as rectangular and circular
L-shaped cross sections, have been used as efficient lateral column sections, can be easily modified for biaxially loaded
force-resisting structures against earthquake and wind loads. walls with various nonplanar section geometries (Hsu 1985,
Such nonplanar walls are subjected to axial compression 1987, 1989).
and combined lateral loads about two orthogonal directions. In fact, the behavior and design of biaxially loaded
Thus, for safety and economy in wall design, complicated columns and walls have been an issue in reinforced concrete
stress and strain distributions over the nonplanar wall cross design, as summarized in detail in Furlong et al. (2004) and
section, caused by a biaxial interaction of axial compression El Fattah et al. (2013). Most previous studies, particularly
and biaxial bending moments, should be taken into account those including straightforward design approaches such as
(refer to Fig. 1). the load contour method and reciprocal load method, have
Various strain compatibility section analysis/design been focused on biaxially loaded rectangular or circular
methods have been developed based on advanced analysis columns with symmetric solid sections. Relatively, research
techniques and enhanced computational power (Lau et al. on the biaxial behavior and design of nonplanar walls with
1993; Ahmad and Weekoon 1995; Chen et al. 2001; de unsymmetrical thin-walled sections is limited. Hsu (1985,
Sousa and Caldas 2005; Rodriguez-Gutierrez and Aristiza- 1987, 1989) studied experimentally and analytically the
bal-Ochoa 1999a,b; Fafitis 2001; Chiorean 2010; Ludovico behavior of biaxially loaded L-shaped, T-shaped, and
et al. 2010; Marmo et al. 2011; Papanikolaou 2012). The C-shaped walls with various axial load levels and weak axis
existing methods have been focused on the development of to strong axis moment ratios, and reported that the same
versatile numerical methods that can accommodate a variety load contour method as that used for rectangular columns
of section geometries (dimensions, cross-sectional shapes, can be used for the design of the L-shaped, T-shaped, and
and reinforcement arrangements) and materials (such as
concrete, steel, and fiber-reinforced polymer). However, ACI Structural Journal, V. 115, No. 1, January 2018.
such numerical methods are not appropriate for general use MS No. S-2016-433.R1, doi: 10.14359/51700917, received January 27. 2017, and
reviewed under Institute publication policies. Copyright © 2018, American Concrete
in practice because they require a dedicated software incor- Institute. All rights reserved, including the making of copies unless permission is
obtained from the copyright proprietors. Pertinent discussion including author’s
porated with pre- and post-processes. closure, if any, will be published ten months from this journal’s date if the discussion
is received within four months of the paper’s print publication.

ACI Structural Journal/January 2018 151


Fig. 1—Stresses and strains at ultimate limit state of nonplanar wall sections subjected to axial compression and combined
biaxial moments.

Fig. 2—Strain-based section analysis method: flowchart for computer coding and stress-strain relationships of concrete and
reinforcing steel.
C-shaped walls. However, recent studies using advanced Finally, the procedure and an example to design a biaxially
analysis methods have shown that the biaxial behavior and loaded T-shaped wall are given.
load contour shape of nonplanar walls might fundamentally
differ from those of columns (Ludovico et al. 2010; Marmo RESEARCH SIGNIFICANCE
et al. 2011; Papanikolaou 2012). For example, as shown in Nonplanar walls with flanged sections are mainly
Fig. 1, because the C-shaped wall is unsymmetrical about the subjected to biaxial moments due to their unsymmetrical
y-axis, its biaxial behavior might be significantly affected by section geometry. The biaxial behaviors of these nonplanar
the ratio of moments about the weak and strong axes. walls fundamentally differ from those of columns with rect-
In this study, a straightforward biaxial design method for angular and circular solid sections. Thus, traditional biaxial
T-shaped walls is proposed by modifying the existing load design methods and design aids, which are mostly provided
contour method. A section analysis program based on the for rectangular and circular columns, cannot be used for
strain compatibility method is developed and its validity nonplanar walls. In this study, characteristics of the biaxial
is verified by comparisons with the existing tests. Then, interaction occurring in T-shaped walls are investigated in
through a parametric study using the developed program, detail, and a modified load-contour method is proposed.
the interaction of biaxial moment strengths at constant axial
loads in monosymmetric T-shaped wall sections is investi- STRAIN COMPATIBILITY SECTION ANALYSIS
gated. Further, based on these investigations, the normal- A strain compatibility section analysis method for
ization method and non-dimensional contour equations of nonplanar walls with arbitrary section geometries is
the biaxial moment strengths, which depend on the direction developed. Figure 2(a) shows the flowchart of the section
of loading and the magnitude of axial loads, are proposed. analysis. First, input data such as geometric and mate-

152 ACI Structural Journal/January 2018


Fig. 3—Comparisons of calculated P-M interaction curves with test results. (Note: 1 in. = 25.4 mm; 1 ksi = 6.9 MPa; 1 kip =
4.448 kN; 1 kip-ft = 1.36 kN∙m.)
rial properties are given. Then, by assuming the angle of where fc and εc are the stress and strain of the concrete, respec-
the neutral axis (–180 degrees ≤ θ ≤ 180 degrees) and the tively; fc′ (>0) is concrete compressive strength; η is residual
depth of the compression zone (c) (refer to Fig. 2(b)), the strength ratio of the concrete considering the post-peak soft-
stress and strain distributions over the wall cross section are ening behavior; εcu (>0) is ultimate compressive strain of the
determined. By numerically integrating the internal tension/ concrete; and εco (>0) is strain corresponding to fc′. The tensile
compression forces and moments of concrete and reinforce- stress of concrete is neglected. The stress-strain relationship
ment elements (that is, δFi, δMxi, and δMyi of element i; refer of steel reinforcements is assumed to follow the elastic-per-
to Fig. 2(b)) over the entire wall section, the axial compres- fectly plastic behavior with Young’s modulus Es (= 200 GPa
sion strength Pn and the bending strengths Mnx and Mny about [29,000 ksi]) and yield strength fy (refer to Fig. 2(d)).
the x- and y-axes, respectively, are estimated. The x- and For verification, in Fig. 3 and Table 1, the P-M interaction
y-axes denote the horizontal and vertical axes, respectively, curves obtained from the proposed method are compared
passing through the geometrical center of the wall section. with the existing wall tests. The wall specimens are biaxi-
The calculation is repeated for increasing θ and c. Through ally-loaded C-shaped, T-shaped, and L-shaped wall speci-
these iterative calculations, the complete three-dimensional mens tested by Hsu (1985, 1987, 1989). Section geometries,
Pn-Mnx-Mny interaction surface of the wall section is obtained. dimensions, and reinforcement details are shown in Fig. 3. The
In addition, through post-processing, the Mnx-Mny contours material properties such as concrete compressive strengths
at a given axial compression strength Pn or the Pn-Mn inter- fc′ and steel yield strengths fy are also shown in Table 1. In
action curve at a given biaxial eccentricity can be calculated. the wall specimens, axial compression loads P are applied
In the section analysis, linear strain distribution is with eccentricities ex and ey in the x- and y-directions,
assumed, as illustrated in Fig. 2(b). For concrete in compres- respectively, from the center of geometry (refer to Fig. 3(a)).
sion, a parabolic-linear stress-strain relationship is used as Through those eccentricities, the biaxial moments Mx (= Pey)
follows (Wight and MacGregor 2012; Rodriguez-Gutierrez and My (= Pex) increasing in linear proportion to the axial
and Aristizabal-Ochoa 1999a,b; refer to Fig. 2(c)) loads P are applied to the walls. The direction angle of
biaxial moments is denoted as ψ, defined as tan–1(Mx/My) or
  2ε  ε  2  tan–1(ey/ex). The ex, ey, and ψ values of the wall specimens
 f c′  c +  c   for − ε co ≤ ε c ≤ 0 are presented in Table 1 (Hsu 1985, 1987, 1989).
  ε co  ε co   Figures 3(b) to 3(e) compare the P-M interaction curves

fc =  (1) obtained from the analyses with the test results. The vertical
 and horizontal axes denote the axial compression load
 − f ′ 1 − (1 − η)  ε c + ε co   for − ε ≤ ε ≤ − ε and bending moment about the x- or y-axis, respectively.
 c  − ε + ε   cu c co
In each figure, two types of theoretical P-M curves are
  cu co 
shown. The P-M curves calculated by the proposed method

ACI Structural Journal/January 2018 153


Table 1—Test variables of nonplanar members considered for verification
Specimen Section Main bars fc′, MPa fy, MPa ex, mm ey, mm ψ, deg Pu, kN
1c 22 No. 3 25.2 357 45.7 76.2 65.6 482
2c 22 No. 3 25.2 357 45.7 69.9 63.1 532
C-shaped:
3c 22 No. 3 25.2 357 45.7 76.2 65.6 459
Type A
4c 22 No. 3 29.2 357 45.7 88.9 69.8 477
5c 22 No. 3 26.9 357 38.1 63.5 65.6 541
8y C-shaped: 18 No. 3 20.4 453 116 78.4 34.1 330
9y Type B 18 No. 3 20.4 453 126 85.6 34.1 294
T-2n 18 No. 3 29.3 448 189 38.1 11.4 216
T-shaped:
T-5n 18 No. 3 33.4 448 147 28.4 10.9 303
Type A1
T-6n 18 No. 3 33.4 532 76.2 14.2 10.6 573
T-1n 18 No. 3 29.3 532 60.2 63.5 46.5 482
T-3n T-shaped: 18 No. 3 29.3 532 53.1 55.4 46.2 619
T-4n Type A1 18 No. 3 33.4 532 58.7 60.2 45.7 528
T-7t 18 No. 3 29.3 532 60.3 63.5 46.5 487
2a L-shaped 14 No. 3 25.9 357 16.3 16.3 45.0 496
4b 14 No. 3 29.0 462 38.9 127 73.0 170
L-shaped
5b 14 No. 3 29.0 462 42.7 140 73.0 156

Notes: 1 ksi = 6.9 MPa; 1 in = 25.4 mm; 1 kip = 4.45 kN.

are denoted as black solid lines, while the theoretical P-M CHARACTERISTICS OF BIAXIAL INTERACTION
curves reported by Hsu (1985, 1987, 1989) are denoted as OF T-SHAPED WALLS
gray dotted lines. The test results are plotted as white circles Biaxial interaction surfaces of nonplanar walls with thin-
and triangles. It is noted that material strengths (fc′ and fy) walled sections can completely differ to those of columns
and moment directions (ψ) used for both theoretical P-M with rectangular and circular solid sections. In this study,
curves slightly differed from those of the tests. The fc′, fy, and T-shaped walls with monosymmetric sections were chosen
ψ values used for the analyses are shown in Fig. 3(b) to 3(e), for a fundamental study to investigate the characteristics
while the actual fc′, fy, and ψ values of each wall specimen of the biaxial interaction in nonplanar walls because their
are presented in Table 1. For example, as shown in Fig. 3(b) section geometry is relatively simple.
and Table 1, the actual fc′ and ψ values of T-shaped wall The three-dimensional Pn-Mnx-Mny interaction surface of a
specimens 1c, 2c, 3c, 4c, and 5c varied slightly between prototype T-shaped wall, obtained by the proposed section
fc′ = 25.2 and 26.9 MPa (3.65 and 3.90 ksi) and between analysis method, is shown in Fig. 4(a). Sectional and material
ψ = 63.1 and 69.8 degrees; however, the theoretical Pn-Mnx properties of the prototype wall are also shown in Fig. 4(a).
and Pn-Mny curves plotted in the figure are those corre- In fact, it is inconvenient to use such 3-D interaction
sponding to fc′ = 24.1 MPa (3.50 ksi) and ψ = 59 degrees, surface for wall design. Thus, in the load contour method, as
which are the same as the values used by Hsu (1989), so that shown in Fig. 4(b), the interaction surface is converted into
the analysis and test results of the five wall specimens can be two-dimensional Mnx-Mny contours, each of which represents
compared in the same figure (refer to Fig. 3(b)). In the esti- the nominal moment strengths about the x- and y-axes at a
mation of the theoretical Pn -Mnx and Pn -Mny curves by the constant axial strength Pn. The Mnx-Mny contours, repre-
proposed method, the characteristic strains of the concrete sented with different line types and weights, correspond to
are taken as εco = 0.002 and εcu = 0.0038 (Rodriguez-Guti- the axial compression ratios of n = 0.0, 0.1, 0.2, 0.3, 0.4,
errez and Aristizabal-Ochoa 1999a,b). The residual and 0.5, where n = Pn/Agfc′. Each Mnx-Mny contour includes
compressive strength of the concrete at εcu = 0.0038 is results of section analyses for neutral axis angles varying in
assumed as 0.85fc′ (that is, η = 0.85; refer to Fig. 2(c) and the range of –180 degrees ≤ θ ≤ +180 degrees. The neutral
Eq. (1)). axis angle, defined along the counter-clockwise direction
As shown in Fig. 3(b) to 3(e), the agreements between the about the positive x-axis, is denoted as θ (= –180 to +180
theoretical Pn -Mnx and Pn -Mny curves of the present study degrees; refer to Fig. 4(a)). The positive and negative angles
and Hsu’s previous studies are reasonable. In addition, in denote the direction of the compression zone. For example,
most specimens, the test strengths represented as the white θ = +90 and –90 degrees indicate that the neutral axes are
circles and triangles are slightly greater than the theoretical parallel to the y-axis in both cases, but the compression
P-M curves by the proposed method, regardless of the types zones exist in the right and left sides of the neutral axis,
of section geometries and the magnitude of axial compres- respectively. For distinction, the Mnx-Mny contours corre-
sion loads. It is noted that slenderness effect is not included sponding to –180 degrees ≤ θ ≤ 0 degrees and 0 degrees ≤ θ
in Fig. 3 because the specimens tested by Hsu were stocky. ≤ +180 degrees are denoted as gray and black lines, respec-
tively. From Fig. 4(b), the characteristics of the biaxial inter-
action of mono-symmetric T-shaped walls can be summa-
rized as follows:

154 ACI Structural Journal/January 2018


Fig. 4—Conversion Pn-Mnx-Mny interaction surface into family of Mnx-Mny contours at constant axial loads. (Note: 1 in. = 25.4
mm; 1 kip-ft = 1.36 kN∙m.)

Fig. 5—Variation of center of compression depending on angles of neutral axis (n = 1 and 0.5).
1. Because the T-shaped wall section is unsymmetrical 2. The shapes of Mnx-Mny contours vary with the magni-
about the y-axis, the Mnx-Mny contours also have unsym- tude of axial compression loads. For example, as shown in
metrical geometries depending on the sign of Mny. This is Fig. 4(c), the Mnx-Mny contour shape resembles a triangle at
attributed to the unsymmetrical section geometry made up n = 0.1, but changes close to a rectangle at n = 0.5. This can
of relatively thin flange and web, as follows. Figure 5(a) be easily understood by comparing the location of the neutral
shows three neutral axis angles of θ = –45, 0, and 45 degrees axis and the center of compressive forces at n = 0.1 with
in the prototype wall under the same axial compression those at n = 0.5 (refer to Fig. 5(a) and 5(b)). Figure 5(b) shows
load of Pn = 0.1Agfc′ (that is, n = 0.1). In each case, the three neutral axis angles of θ = –45, 0, and 45 degrees at the
compression zone is represented as the shaded area and the same axial compression ratio of n = 0.5. The neutral axes
center of internal compression forces are denoted as black at n = 0.5 are above the web due to the increased axial
dot. In addition, the distances between the center of compression load. Thus, when the neutral axis angle
compressive forces to the x- and y-axes are denoted as yo increases from θ = –45 to +45 degrees, Mnx remains almost
and xo, respectively. If the neutral axis angle decreases unchanged because yo does not change significantly. On the
from θ = 0 to –45 degrees (refer to Fig. 5(a2)), the center of other hand, Mny changes significantly from negative to posi-
compressive forces does not change significantly, and conse- tive as the center of compressive forces moves from the left
quently Mnx and M ny remain almost unchanged. Thus, as of the y-axis at θ = –45 degrees to the right at θ = +45 degrees.
shown in Fig. 4(c), the points corresponding to θ = 0 and 3. In T-shaped wall sections, the geometrical center does
–45 degrees in the Mnx-Mny contour at n = 0.1 are almost in the not coincide with the shear center. Thus, the neutral axis angle
same location. In contrast, if the neutral axis angle increases θ differs from the direction of biaxial moments ψ, defined as
from θ = 0 to +45 degrees so that the neutral axis intersects tan–1(Mnx/Mny). For example, as shown in Fig. 4(c), although
both the flange and web (refer to Fig. 5(a3)), the center of the neutral axis is parallel to the x axis (that is θ = 0 degrees),
compressive forces moves upward close to the x-axis and to the moment about the y-axis, Mny, cannot be zero (ψ ≠ 0
the right crossing the y-axis. In other words, Mnx decreases degrees). This indicates that the direction of bending differs
and Mny changes from negative to positive. Thus, in the from the direction of lateral loading. Consequently, T-shaped
Mnx-Mny contour at n = 0.1 of Fig. 4(c), the point corre- walls resist biaxial bending accompanied by torsion and
sponding to θ = +45 degrees, denoted as a triangle, is twisting deformation. In Fig. 6, although the lateral load acts
displaced to the first quadrant. along the y-axis and the wall is thus subjected to uniaxial

ACI Structural Journal/January 2018 155


Fig. 6—Twisting of cantilever T-shaped wall due to lateral load acting along y-axis.

Fig. 7—Mapping of Mnx-Mny contour into non-dimensional mx-my contour through line transformation.
bending about the x-axis, a twisting deformation occurs in portions at 0 degrees ≤ θ ≤ +90 degrees and –90 degrees ≤
the wall. θ ≤ 0 degrees of Mnx-Mny contour are mapped to the non-di-
mensional mx-my contours at the first and forth quadrants,
CONTOUR EQUATIONS FOR T-SHAPED WALLS respectively, through a linear transformation defined as
Normalization of contours follows (refer to Fig. 7(b))
The investigation results of the previous section show
that, to address the characteristics of biaxial interaction M nx
mx = for − 90 deg ≤ θ ≤ 90 deg (2)
in T-shaped walls, the effects of the direction of bending M nx 0
moments and the magnitude of axial compression load on
the contour geometry should be taken into account.
Figure 7 shows the normalization process for unsym-  M ny − M ny 0
 for 0 deg ≤ θ ≤ 90 deg [or M ny − M ny 0 ≥ 0]
metrical Mnx-Mny contours of T-shaped walls. In the figure, M − M ny 0
my =  ny 90
points on the Mnx-Mny contour corresponding to θ = 0, 90,  M ny − M ny 0
for − 90 deg ≤ θ ≤ 0 deg [or M ny − M ny 0 < 0]
and –90 degrees are denoted as white and gray circles for  M ny′ 90 − M ny 0

distinction. Mnx0 and Mny0 are the moment strengths about
the x- and y-axes at θ = 0 degrees, respectively, and Mny90 (3)
and M ′ny90 are the moment strengths about the y-axis at θ
= 90 and –90 degrees, respectively. Because T-shaped wall As shown in Fig. 7, through the normalization by the
sections are symmetrical about the x-axis, the contour shape proposed linear transformation, the unsymmetrical char-
is the same for positive and negative Mnx. Thus, the contour acteristics of the Mnx-Mny contours, depending on design
of only positive Mnx (that is, –90 degrees ≤ θ ≤ +90 degrees) parameters, are significantly alleviated in the mx-my contours.
is considered for simplicity. For distinction, the contours for
–90 degrees ≤ θ ≤ 0 degrees and 0 degrees ≤ θ ≤ 90 degrees Elliptic contour equations
are denoted as solid and dashed lines, respectively. As In the load contour method, the non-dimensional mx-my
shown in Fig. 7(a), the Mnx-Mny contours for –90 degrees ≤ contour is basically defined as the following elliptical equation.
θ ≤ 0 degrees and 0 degrees ≤ θ ≤ 90 degrees significantly
differ in size because the neutral axis angle θ differs from (mx)α + (my)α = 1 (4)
the biaxial moment angle ψ (compare the white and shaded
areas). To relieve the asymmetry in the contour shape, the where α is the factor affecting the shape of the contour varying
with design parameters such as material strength, section

156 ACI Structural Journal/January 2018


Fig. 8—Mnx-Mny and mx-my contours for prototype T-shaped wall sections varying with axial load ratios (n = 1 to 0.5). (Note:
1 in. = 25.4 mm; 1 ksi = 6.9 MPa; 1 kip-ft = 1.36 kN∙m.)
geometry, axial compression load, and biaxial moment (58,000 psi), respectively. In addition, it is assumed that
direction. According to Bresler (1960) and the PCA Notes on longitudinal reinforcements are uniformly distributed over
ACI 318-11 (Portland Cement Association 2013), the values the entire wall section with a reinforcement ratio of ρ = 0.01.
of α in the rectangular and circular columns vary between The Mnx-Mny and mx-my contours of five T-shaped wall
1.15 and 1.55. Hsu (1985, 1987, 1989) suggested α = 1.5 sections are shown in Fig. 8(b) to 8(f). In each figure, the
for L-shaped, T-shaped, and C-shaped walls. However, it is contours are estimated as follows. By using the proposed
inappropriate to use such a single value of α for T-shaped strain compatibility section analysis method, the three-
walls, considering that the unsymmetrical contour shape is dimensional Pn-Mnx-Mny interaction surfaces of T-shaped
significantly affected by the sign of moments (that is, posi- walls are obtained and the Mnx-Mny contours corresponding
tive and negative Mny) and the magnitude of axial loads (n). to n = 0.1 to 0.5, shown in Fig. 8(b1) to 8(f1), are then plotted.
Thus, α values appropriate for T-shaped walls are investi- The Mnx-Mny contours are converted into non-dimensional
gated through a numerical parametric study as follows. mx-my contours presented in Fig. 8(b2) to 8(f2) through the
Figure 8(a) shows the cross sections of five T-shaped walls linear transformation following Eq. (2) and (3). In Fig. 8, for
used for the parametric study. In the figure, the symbols t, b, distinction, the contours for no or low axial compression
and h denote the wall thickness (= 200 mm [8 in.]), web ratios of n = 0, 0.1, and 0.2 are denoted as thin solid lines
length, and half the flange length, respectively. By varying connecting solid diamonds, circles, and triangles, respec-
h and/or b between 500 and 4000 mm (19.7 and 157.5 in.), tively, while the contours for relatively high axial compres-
the shape factor γ, which is defined as the ratio of the half sion ratios of n = 0.3, 0.4, and 0.5 are denoted as thin dashed
flange length to the web length (= b/h), is gradually increased lines connecting voided diamonds, circles, and triangles,
from 0.25 to 0.5, 1.0, 1.5, and 2.0. In the parametric study, respectively. It should be noted that the maximum design
the concrete compressive strength and steel yield strength axial compression strength is ϕPn,max = ϕ · 0.8(0.85fc′[Ag –
are taken as fc′ = 30 MPa (4350 psi) and fy = 400 MPa As] + fy As) and, when substituting ϕ = 0.65, ϕPn,max is less

ACI Structural Journal/January 2018 157


Fig. 9—Effects of other design variables such as reinforcement ratio, reinforcement arrangement, concrete strength, and steel
yield strength on normalized mx-my contour shapes. (Note: 1 in. = 25.4 mm; 1 ksi = 6.9 MPa.)
than 0.5Ag fc′. Thus, in this study, the axial compression ratio significantly modified in the region of –90 degrees ≤ θ ≤ 0
is limited to 0 ≤ n ≤ 0.5. degrees through the linear transformation (refer to Fig. 7).
As shown in Fig. 8(b2) through 8(f2), the mx-my contours of
T-shaped wall sections differ in their geometries depending Effects of other design parameters
on the neutral axis angle θ. In the region of 0 deg ≤ θ ≤ +90 According to the PCA Notes on ACI 318-11 (Portland
degrees (that is, in the first quadrant), the mx-my contours Cement Association 2013), the mx-my contours can also be
expand as the axial compression ratio increases from n = 0 affected by other parameters such as reinforcement ratio (ρ),
to 0.5. This indicates that a greater α can be used for high reinforcement arrangement (distributed and end-concen-
axial loads. However, in the region of –90 degrees ≤ θ ≤ trated), concrete strength (fc′), and steel yield strength (fy). The
0 degrees (that is, in the fourth quadrant), the correlation effects of such design parameters are investigated through
between the axial compression ratio n and the shape of the additional analyses for the prototype wall with a shape factor
mx-my contours (or α) is relatively weak. In both regions, γ = 1.0. Figure 9(a) compares the mx-my contours for ρ = 0.005,
the effects of the shape factor γ (= b/h = 0.25 to 2.0) on the 0.01, and 0.02. In the cases of ρ = 0.005 and 0.01, the differ-
mx-my contour shape are also limited. Thus, based on these ence in the contour shapes is not significant (refer to Fig. 9(a1)
results, the α values for biaxially-loaded T-shaped walls are and 9(a2)). However, as the reinforcement ratio increases to
proposed as follows ρ = 0.02, the mx-my contours expand in the region of 0 degrees ≤
θ ≤ +90 degrees (that is, in the first quadrant; refer to
3n + 1.05 for 0 deg ≤ θ ≤ 90 deg Fig. 9(a3)). This indicates that the α values defined in Eq. (5)
α= (5) are conservative for ρ = 0.005 to 0.02.
2.0 for − 90 deg ≤ θ ≤ 0 deg
Figure 9(b) shows the mx-my contours of the prototype
In Fig. 8(b2) to 8(f2), the proposed mx-my contours, repre- T-shaped wall section with end-concentrated reinforce-
sented as thick solid and dashed lines, are compared with ment arrangement. As shown in Fig. 9(b), the reinforcement
the analysis results (that is, thin gray lines with solid/voided ratio at the ends of the flange and web is increased to ρe =
diamonds, circles, or triangles). The proposed contours are 0.023, but the overall reinforcement ratio is maintained as ρ
obtained by substituting the α values defined in Eq. (5) into = 0.01. When compared with Fig. 9(a2), changes in the
Eq. (4). For clarity, the proposed contours in the region of contour shapes due to distributed and end-concentrated
0 degrees ≤ θ ≤ +90 degrees are plotted only for n = 0 (that reinforcement arrangements are almost negligible. Figures 9(c)
is, α = 1.05) and 0.5 (that is, α = 3.05). In most cases, the and 9(d) show the mx-my contours for an increased concrete
proposed α values are conservative compared to the analysis strength fc′ = 40 MPa (5800 psi) and for an increased steel
results. However, as presented in Fig. 8(e2) and 8(f2), the yield strength fy = 600 MPa (87 ksi), respectively. Compared
proposed α = 2.0 in the region of –90 degrees ≤ θ ≤ 0 degrees with the contours of the prototype wall shown in Fig. 9(a2),
is slightly overestimated for n = 0. Such overestimation is changes in the contour shapes due to the increased concrete
limited in the Mnx-Mny contour, when considering the fact strength and steel yield strength are also limited. In summary,
that the mx-my contours at low axial compression ratio are Fig. 9(b) to 9(d) show that the effects of the reinforcement

158 ACI Structural Journal/January 2018


Fig. 10—Proposed Mnx-Mny contour for biaxial design of T-shaped walls.
arrangement and material strengths on the contour shapes (–1475 kip-ft), respectively. The biaxial design of the
are limited. T-shaped wall is performed according to the following
procedures:
DESIGN APPLICATION 1. Assume a longitudinal reinforcement ratio ρ of the wall
For biaxial design of T-shaped walls, the ϕMnx-ϕMny section.
contour at a constant axial compression strength ϕPn is 2. By performing section analyses for the neutral axis
presented in Fig. 10 (ϕ = strength reduction factor). By angles θ = –90, 0, and +90 deg, two-dimensional ϕPn-ϕMnx
substituting Eq. (2) and (3) into Eq. (4), design elliptical and ϕPn-ϕMny interaction curves are obtained. Then, the
contour equations can be defined as follows. characteristic moment strengths ϕMnx0, ϕMny0, ϕMny90, and
For 0 degrees ≤ θ ≤ +90 degrees (or Muy – ϕMny0 ≥ 0): ϕM n′ y90 corresponding to the factored design axial load Pu are
estimated from the interaction curves (refer to Fig. 11(b)).
 M ux 
α
 M uy − φM ny 0 
α
3. The α is determined from Eq. (5) using n = Pu/(Agfc′).
α α
m +m =
ux uy  +  ≤ 1.0 (6a) Then, the biaxial design contour of the wall section is drawn
 φM nx 0   φM ny 90 − φM ny 0  in accordance with Eq. (6a) and (6b) and Fig. 10.
4. Check whether the biaxial moments Mux and Muy are
For –90 degrees ≤ θ ≤ 0 degrees (or Muy – ϕMny0 < 0):
within the design contour. If not, try Steps 1 through 4 again,
α α changing the reinforcement ratio ρ.
 M ux   M uy − φM ny 0  Design results of the biaxially-loaded T-shaped wall are
muxα + muyα =   +  ≤ 1.0 (6b)
 φM nx 0   φM ny′ 0 − φM ny 0  shown in Fig. 11 and Table 2. As a first trial, the reinforce-
ment ratio is taken as ρ = 0.008. The ϕPn-ϕMnx and ϕPn-ϕMny
In these equations, Mux and Muy are the design moments interaction curves for θ = –90, 0, and 90 degrees are shown
about the x- and y-axes, respectively, acting on T-shaped in Fig. 11(b). The strength reduction factor ϕ and maximum
walls, and all nominal moment strengths are multiplied axial compression strength Pn,max are determined in accor-
by the ϕ factor for design. In addition, the α factor corre- dance with ACI 318-14. The characteristic moment strengths
sponding to a design axial load Pu (that is, n = Pu/[Agfc′]) is ϕMnx0 (= 7571 kN∙m [5584 kip-ft]), ϕMny0 (= –1682 kN∙m
determined from Eq. (5). To draw a complete design contour [–1241 kip-ft]), ϕMny90 (= 5606 kN∙m [4135 kip-ft]), and
for T-shaped walls, characteristic moment strengths such ϕM n′ y90 (= –2181 kN∙m [–1609 kip-ft]) corresponding to
as Mnx0, Mny0, Mny90, and M n′ y90 at a given axial load Pu (= the factored design axial load Pu (= 2832 kN [637 kip]) are
ϕPn) should be estimated first, as denoted with white and then estimated, as shown in Fig. 11(b) and Table 2. Note
gray circles in Fig. 10. Such characteristic design moment that the moment about the y-axis, Muy (= –2000 kN∙m
strengths can be determined without significant computa- [–1475 kip-ft]), is less than ϕMny0 (= –1682 kN∙m [–1241
tional efforts because the neutral axes are parallel to the x- kip-ft]) and, thus, the neutral axis angle exists in the region
or y-axis. of –90 degrees ≤ θ ≤ 0 degrees. In addition, the α values
A design example of a biaxially loaded T-shaped wall corresponding to the axial compression ratio n = Pu/(Agfc′) =
using the proposed method is presented as follows. Figure 11(a) 0.1 are determined as 2.0 and 1.45 in the regions of –90
shows the section properties and material strengths of the degrees ≤ θ ≤ 0 degrees and 0 degrees ≤ θ ≤ +90 deg, respec-
wall. The factored design axial load, moment about the x-axis, tively. Figure 11(c) shows the mx-my contour for the biaxial
and moment about the y-axis are Pu = 2832 kN (637 kip), design of the T-shaped wall, plotted in the non-dimensional
Mux = 5500 kN·m (4057 kip-ft), and Muy = –2000 kN∙m plane. As shown in Fig. 11(c), mux = 0.726 and muy = 0.637,

ACI Structural Journal/January 2018 159


Table 2—Summary of biaxial design of T-shaped wall
Biaxial strengths at θ = 0, 90, and –90
Design loads deg Non-dimensional load contour
Mux, Muy, ϕMnx0, ϕMny0, ϕMny90, ϕM n′ y90, α α
Pu, kN kN·m kN·m kN·m kN·m kN·m kN·m n (–90 deg ≤ θ ≤ 0 deg) (0 deg ≤ θ ≤ +90 deg) mux muy
2832 5500 –2000 7571 –1682 5606 –2181 0.1 2.0 1.35 0.726 0.637

Notes: 1 kip-ft = 1.36 kN-m; 1 kip = 4.45 kN

Fig. 11—Flexure-compression design of biaxially loaded T-shaped wall using proposed load contour method. (Note: 1 in. =
25.4 mm; 1 ksi = 6.9 MPa; 1 kip = 4.45 kN; 1 kip-ft = 1.36 kN∙m.)
represented as a black dot in the fourth quadrant, satisfy SUMMARY AND CONCLUSIONS
Eq. (6b): mux2 + muy2 = 0.7262 + 0.6372 = 0.933 < 1.0. In this study, the load contour method for biaxially loaded
In this study, a modified load contour method for biaxial T-shaped walls is developed. Through the theoretical para-
design of T-shaped walls is proposed. Because the elliptical metric study using the strain-compatibility section analysis,
contour equations and exponent α are proposed from the the characteristics of the biaxial interaction of T-shaped
theoretical investigation for limited ranges of design param- walls have been investigated. Based on the results, non-
eters, the application of the proposed method needs to be dimensional contour equations are proposed. In addition, the
limited accordingly. Thus, the proposed method is appli- wall design procedure using the proposed contour equations
cable only to T-shaped walls with monosymmetric sections is given. Conclusions are as follows:
where the shape factor representing the ratio of half flange 1. In T-shaped walls, the Mnx-Mny contours at constant axial
length-to-web length should be between γ = 0.25 and 2.0. compression loads are unsymmetrical in shape and signifi-
In addition, the results of the parametric study show that, cantly affected by neutral axis angles (that is, directions of
at least within ρ = 0.005 to 0.02, fc′ = 30 to 40 MPa (4350 loading) and axial load levels. The unsymmetrical contour
to 5800 psi), and fy = 400 to 600 MPa (58 to 87 ksi), the shape is attributed to the mono-symmetrical section geom-
proposed α and elliptical contour equations are conservative. etry made up of thin flange and web. Because the geometric
It is noted that the wall length-to-thickness ratio of the flange centroid does not coincide with the shear center, T-shaped
and web varies between 2.0 to 9.75, as shown in Fig. 8(a). walls resist biaxial bending accompanied by torsion. Overall
Thus, conservatively, it is recommended that the length-to- the Mnx-Mny contour shape is close to a triangle at low axial
thickness ratio of both the flange and web in T-shaped walls compression load levels (n = 0.0 to 0.2), but turns close to
should be 3 or greater.

160 ACI Structural Journal/January 2018


a rectangle or square at high axial compression load levels Ahmad, S. H., and Weekoon, S. L., 1995, “Model for Behavior of
Slender Reinforced Concrete Columns Under Biaxial Bending,” ACI Struc-
(n = 0.3 to 0.5). tural Journal, V. 92, No. 2, Mar.-Apr., pp. 188-198.
2. To eliminate effects of torsion, the Mnx-Mny contour of Bresler, B., 1960, “Design Criteria for Reinforced Concrete Columns
T-shaped walls is converted into the non-dimensional mx-my under Axial Load and Biaxial Bending,” ACI Journal Proceedings, V. 57,
No. 5, May, pp. 481-490.
contour through a linear transformation. The mx-my contours Cedolin, L.; Cusatis, G.; Eccheli, S.; and Roveda, M., 2008, “Capacity of
change in shape depending on the axial compression ratio Rectangular Cross Sections under Biaxially Eccentric Loads,” ACI Struc-
n and neutral axis angle θ. As the axial compression ratio tural Journal, V. 105, No. 2, Mar.-Apr., pp. 215-224.
Chen, S. F.; Teng, J. G.; and Chan, S. L., 2001, “Design of Biaxially
n increases, the mx-my contours expand in the region of Loaded Short Composite Columns of Arbitrary Section,” Journal of
0  degrees ≤ θ ≤ +90 degrees. However, in the region of Structural Engineering, ASCE, V. 127, No. 6, pp. 678-685. doi: 10.1061/
–90  degrees ≤ θ ≤ 0 degrees, the correlation between the (ASCE)0733-9445(2001)127:6(678)
Chiorean, C. G., 2010, “Computerised Interaction Diagrams and
axial compression ratio n and the mx-my contour shape is Moment Capacity Contours for Composite Steel-Concrete Cross-Sections,”
relatively weak. Other design parameters such as the shape Engineering Structures, V. 32, No. 11, pp. 3734-3757. doi: 10.1016/j.
factor γ (= ratio of flange length-to-web length), reinforce- engstruct.2010.08.019
de Sousa, J. B. Jr., and Caldas, R. B., 2005, “Numerical Analysis of
ment ratio and arrangement, concrete compressive strength, Composite Steel-Concrete Columns of Arbitrary Cross Section,” Journal of
and steel yield strength do not have significant impacts on Structural Engineering, ASCE, V. 131, No. 11, pp. 1721-1730. doi: 10.1061/
the mx-my contour shape. (ASCE)0733-9445(2005)131:11(1721)
El Fattah, A. M.; Rasheed, H.; and Esmaeily, A., 2013, “LRFD Soft-
3. Elliptical load contour equations are proposed based on ware for Design and Actual Ultimate Capacity of Confined Rectangular
the parametric study on prototype T-shaped walls. Because Columns,” Report No. K-TRAN: KSU-11-3, Kansas Department of Trans-
the section geometry and contour shape are unsymmetrical portation, Topeka, KS, 220 pp.
Fafitis, A., 2001, “Interaction Surfaces of Reinforced-Concrete Sections
about the y-axis, different equations are used depending on in Biaxial Bending,” Journal of Structural Engineering, ASCE, V. 127,
the neutral axis angle θ or (the sign and magnitude of the No. 7, pp. 840-846. doi: 10.1061/(ASCE)0733-9445(2001)127:7(840)
y-axis moment Mny). The parametric study shows that the Furlong, R. W.; Hsu, C.-T. T.; and Mirza, S. A., 2004, “Analysis and
Design of Concrete Columns for Biaxial Bending—Overview,” ACI Struc-
proposed contour equations are valid for n = 0.0 to 0.5, γ = tural Journal, V. 101, No. 3, May-June, pp. 413-423.
0.25 to 2.0, ρ = 0.005 to 0.02, fc′ = 30 to 40 MPa (4350 to Hsu, C.-T. T., 1985, “Biaxially Loaded L-Shaped Reinforced Concrete
5800 psi), and fy = 400 to 600 MPa (58 to 87 ksi). Columns,” Journal of Structural Engineering, ASCE, V. 111, No. 12,
pp. 2576-2595. doi: (errata, 1988, V. 114 No. 11, p. 2629)10.1061/
4. The elliptical contour equations can be used for prelim- (ASCE)0733-9445(1985)111:12(2576)
inary design of biaxially loaded T-shaped walls in selecting Hsu, C.-T. T., 1987, “Channel-Shaped Reinforced Concrete Compres-
a trial section, because the load contour method does not sion Members under Biaxial Bending,” ACI Structural Journal, V. 84, No. 3,
May-June, pp. 201-211.
require significant computational efforts nor dedicated soft- Hsu, C.-T. T., 1989, “T-Shaped Reinforced Concrete Members under
ware. The advantage of the proposed design method is that Biaxial Bending and Axial Compression,” ACI Structural Journal, V. 86,
the biaxial design of nonplanar walls can be performed by No. 4, July-Aug., pp. 460-468.
Lau, C. Y.; Chan, S. L.; and So, A. K. W., 1993, “Biaxial Bending Design
using only uniaxial P-M interaction curve about each axis of Arbitrarily Shaped Reinforced Concrete Column,” ACI Structural
(that is, ϕPn-ϕMnx and/or ϕPn-ϕMny interaction curves for θ = Journal, V. 90, No. 3, May-June, pp. 269-278.
–90, 0, and 90 degrees). Ludovico, M. D.; Lignola, G. P.; Prota, A.; and Cosenza, E., 2010,
“Nonlinear Analysis of Cross Sections under Axial Load and Biaxial
Bending,” ACI Structural Journal, V. 107, No. 4, July-Aug., pp. 390-399.
AUTHOR BIOS MacGregor, J. G., 1973, “Simple Design Procedures for Concrete
Tae-Sung Eom is an Associate Professor in the Department of Architec- Columns,” Introductory Report, Symposium on Design and Safety of Rein-
tural Engineering at Dankook University, Yongin, South Korea. He received forced Concrete Compression Members, Reports of the Working Commis-
his BE, MS, and PhD in architectural engineering from Seoul National sions, V. 15, International Association of Bridge and Structural Engineering,
University, Seoul, South Korea. His research interests include nonlinear Zurich, Switzerland, Apr., pp. 23-49.
analysis and seismic design of reinforced concrete structures. Marmo, F.; Serpieri, R.; and Rosati, L., 2011, “Ultimate Strength
Analysis of Prestressed Reinforced Concrete Sections under Axial Force
Hye-Sung Nam is a Graduate Student in the Department of Architectural and Biaxial Bending,” Computers & Structures, V. 89, No. 1-2, pp. 91-108.
Engineering at Dankook University, where she received her BE in architec- doi: 10.1016/j.compstruc.2010.08.005
tural engineering. Papanikolaou, V. K., 2012, “Analysis of Arbitrary Composite Sections
in Biaxial Bending and Axial Load,” Computers & Structures, V. 98-99,
Su-Min Kang is an Assistant Professor in the Department of Architectural pp. 33-54. doi: 10.1016/j.compstruc.2012.02.004
Engineering at Chungbuk National University, Cheongju, South Korea. Parme, A. L.; Nieves, J. M.; and Gouwens, A., 1966, “Capacity of Rein-
He received his BE, MS, and PhD in architectural engineering from Seoul forced Rectangular Columns Subjected to Biaxial Bending,” ACI Journal
National University. His research interests include earthquake design of Proceedings, V. 63, No. 9, Sept., pp. 911-923.
reinforced concrete structures. Portland Cement Association, 1966, “Capacity of Reinforced Rectan-
gular Columns Subject to Biaxial Bending,” Publication EB011D, Portland
Cement Association, Skokie, IL.
ACKNOWLEDGMENTS Portland Cement Association, 2013, “Notes on ACI 318-11 Building
This research was conducted with research funds from the Mid-
Code Requirements for Structural Concrete with Design Applications,”
Career Research Program of National Research Foundation of Korea
Publication EB716, Portland Cement Association, Skokie, IL.
(NRF-2015R1A2A2A01003777) and from the Infrastructure and Transporta-
Rodriguez-Gutierrez, J. A., and Aristizabal-Ochoa, 1999a, “Biaxial
tion Technology Promotion Research Program (Code 17CTAP-C129746-01),
Interaction Diagrams for Short RC Columns of Any Cross Section,”
funded by the Ministry of Land, Infrastructure, and Transportation of Korea.
Journal of Structural Engineering, ASCE, V. 125, No. 6, pp. 672-683. doi:
10.1061/(ASCE)0733-9445(1999)125:6(672)
REFERENCES Rodriguez-Gutierrez, J. A., and Aristizabal-Ochoa, J. D., 1999b,
ACI Committee 318, 2011, “Building Code Requirements for Structural “Partially and Fully Prestressed Concrete Sections under Biaxial Bending
Concrete (ACI 318-11) and Commentary,” American Concrete Institute, and Axial Load,” ACI Structural Journal, V. 97, No. 4, Mar.-Apr.,
Farmington Hills, MI, 503 pp. pp. 553-563.
Wight, J. K., and MacGregor, J. G., 2012, Reinforced Concrete:
Mechanics and Design, sixth edition, Pearson Education, Inc., Upper
Saddle River, NJ.

ACI Structural Journal/January 2018 161


APPLY FOR
ACI Foundation
Research Grants
The ACI Foundation annually funds research projects with grants
up to $50,000. The Foundation seeks to advance the concrete
industry through the funding of concrete research projects that
further the knowledge and sustainability of concrete materials,
construction, and structures.

• Topics are encouraged from all areas of concrete research;


• A letter of support of the research concept by an ACI Technical
Committee is required;
• Industry partnering and project cost sharing are encouraged;
• Principal investigators must follow the ACI Foundation’s
published Concrete Research Council Grant Proposal Guide.

The Foundation will begin accepting proposals at the end of


August through December 1. Applications are submitted online
at concreteresearchcouncil.org.
ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 115-S13

Eccentric Punching Shear of Waffle Slab


by Ahmed Faleh Al-Bayati, Lau Teck Leong, and Leslie A. Clark

This paper presents test results of six 1/10th-scale micro-concrete


waffle slabs subjected to internal punching shear in the presence
of moment transfer. Although the observed punching failure mech-
anism of a waffle slab was found to be very similar to that of a
flat slab, the shear capacity is relatively reduced because some of
the potential failure surface is lost when it extends into the waffle
section. Because the current codes of practice do not consider the
punching shear mechanism of waffle slabs, the comparisons with
the test results revealed that BS8110 and EC2 overestimated the
punching capacities of waffle slabs, whereas ACI’s prediction was
conservative. A prediction model based on the upper-bound plas-
ticity theory was proposed to predict the failure load, which gave
good agreement with the tests.

Keywords: design; internal punching shear; moment transfer; plasticity;


reinforced concrete; waffle slab.

INTRODUCTION
Waffle slabs have been widely used owing to their economic
benefits. It consists of a grid of ribs, distributed in orthogonal
directions, regularly spaced, and topped by a thin slab (refer
to Fig. 1). The overall slab is then supported by columns at
various positions, at which solid sections are introduced to
allow for load transfer from the slab to the columns.
Similar to a flat slab, a waffle slab can develop a local shear
failure known as “punching shear failure” (refer to Fig. 2).
At failure, a solid revolution of concrete (“I”) surrounded by
the inclined shear cracks separates normally from the slab, Fig. 1—Waffle slab.
leaving the rest of the slab (“II”) remaining uncracked.1
flat slabs. So far, there has been a very limited amount of
However, despite of being widely used, a small amount of
work conducted to investigate the punching mechanism
research2-7 has been conducted to investigate the punching
of waffle slabs. As a result, it is not clear how to apply the
behavior of waffle slabs. As a result, the shear design proce-
current design methods that are derived from tests on flat
dures of waffle slabs are not included in the current design
slab to design against the punching failure of waffle slabs.
codes.8-10 Therefore, it is not clear how to apply (if neces-
Therefore, there is a need to conduct experimental inves-
sary) the codes’ design clauses for flat slabs to waffle slabs
tigation to aid the understanding of the punching failure
because when the solid section are very wide or top slabs
mechanism of waffle slabs and hence to develop a relevant
are sufficiently thick, the punching failure surface could
prediction model.
form within the solid section (refer to Fig. 2(b) and (c)),
but when the solid section is narrower, the punching failure
EXPERIMENTAL INVESTIGATIONS
surface could pass through the reduced depth section (refer
Specimen’s details
to Fig. 2(a)).11 As a result, a smaller shear failure surface
Tests were performed on six (1/10th-scaled) micro-
could be mobilized, which would consequently lead to a
concrete waffle slab specimens to simulate the internal
lower punching shear capacity.
column-waffle slab connections in the presence of moment
A prediction model based on the upper-bound solution
transfer. Variables considered were the size of solid section
of plasticity theory was proposed to predict the eccentric
and the column’s eccentricity (refer to Table 1).
punching shear of the waffle slab. The proposed model
develops the concentric punching shear model11 to predict
the eccentric punching shear of waffle slabs. ACI Structural Journal, V. 115, No. 1, January 2018.
MS No. S-2016-434.R1, doi: 10.14359/51701090, was received January 19, 2017,
and reviewed under Institute publication policies. Copyright © 2018, American
RESEARCH SIGNIFICANCE Concrete Institute. All rights reserved, including the making of copies unless
Despite the popularity of waffle slabs, research to date permission is obtained from the copyright proprietors. Pertinent discussion including
author’s closure, if any, will be published ten months from this journal’s date if the
has focused on the punching shear failure mechanism of discussion is received within four months of the paper’s print publication.

ACI Structural Journal/January 2018 163


The ratios of column eccentricity to column size, e/cx, casting position of the prototype structures. The moment
considered were 0.5, 1.0, and 1.5. All slab specimens had transferring column(s) were simulated with a steel column
an overall depth of 70 mm (2.8 in.) and a top slab of 20 mm stub with overhang bolted to the specimens (refer to Fig. 4).
(0.8 in.) thick (refer to Fig. 3). These geometrical proper- Although such an arrangement is different from the actual
ties were chosen in accordance with Clause 5.3 in EC2.9 All details of a column slab connection, it is believed that as long
specimens were cast in an upright position to simulate the as the punching shear failure occurs outside the column, the
punching resistance would not be affected.12

Concrete
All specimens were cast from micro-concrete having
a maximum aggregate size of 2 mm (0.08 in.), water-
cement ratio (w/c) of 0.56, and aggregate/cement ratio of
1.6. All the aggregates used were sieved and remixed to the
required grading13 (refer to Table 2). The concrete compres-
sion strengths were obtained from 50 mm (2 in.) cubes cast
simultaneously with slab specimens. However, for calculating
the shear capacities, the measured 50 mm (2 in.) cube concrete
compressive strengths have been modified to represent the
intended 15 mm (0.6 in.) cube compressive strength in accor-
dance to Endersbee’s equation14 to allow for size effects.

−0.106
 15 
f cu15 = f cu 50   (1)
 50 

where both fcu15 and fcu50 are in MPa.

Reinforcements
Due to the difficulties in obtaining small-diameter ribbed
steel bars, reinforcements were fabricated from plain steel
bars of 3.4 mm (0.13 in.) diameter, with an average yield
strength of 440 MPa (63.8 ksi) (refer to Fig. 5). All slab
specimens tested in this research have been reinforced with
a sufficient amount of steel bars with tension bars bent up
at both ends to avoid premature bond failure. Confidence
in the use of small-scaled specimens cast with plain scaled
steel bars further derived from the test results (Fig. 6),
indicating all slab specimens failed in punching with the
manner of their load-deflection curves similar to flat slabs
cast with high-tension reinforcement ratio reported in the
literatures.15 Tension reinforcements were placed at 11 mm
(0.4 in.) spacing across the top slab regions, while all ribs
were doubly reinforced. And 5 mm (0.2 in.) covers to rein-
forcements were maintained in all specimens.

Fig. 2—Punching shear mechanism of waffle slab.

Table 1—Details of waffle slab specimens


Size of solid section, c, mm d, mm d1, mm d2, mm fcu50, N/mm2 Column eccentricity e,
Slab No. mm (in.) (in.) (in.) (in.) (in.) ρ, % ρct, % ρcb, % (psi) mm (in.)
IWSM1 290 x 290 (11.4 x 11.4) 100 (4) 62 (2.44) 63 (2.48) 63 (2.48) 1.29 1.27 0.6 29.5 (4278) 100 (4)
IWSM2 470 x 470 (18.5 x 18.5) 100 (4) 62 (2.44) 63 (2.48) 63 (2.48) 1.29 1.27 0.6 34.7 (5033) 100 (4)
IWSM3 200 x 200 (7.9 x 7.9) 100 (4) 62 (2.44) 63 (2.48) 63 (2.48) 1.29 1.27 0.6 34.6 (5018) 100 (4)
IWSM4 250 x 250 (9.8 x 9.8) 100 (4) 62 (2.44) 63 (2.48) 63 (2.48) 1.29 1.27 0.6 35.6 (5163) 100 (4)
IWSM5 250 x 250 (9.8 x 9.8) 100 (4) 62 (2.44) 63 (2.48) 63 (2.48) 1.29 1.27 0.6 32.5 (4714) 50 (2)
IWSM6 250 x 250 (9.8 x 9.8) 100 (4) 62 (2.44) 63 (2.48) 63 (2.48) 1.29 1.27 0.6 31.7 (4598) 150 (6)
Notes: d is average effective depth of orthogonal tension reinforcement; d1 is effective depth of tension reinforcement in bending direction; d2 is effective depth of compression
reinforcement

164 ACI Structural Journal/January 2018


Fig. 3—Waffle slab specimens.

Fig. 4—Test setup.


Test procedures Table 2—Sieve analysis of aggregates10
All specimens were supported on rollers at four edges in Sieve size Mass retained, g % retained % passing
an inverse manner and loaded eccentrically by means of a
2.0 0.00 0.00 100.00
hand-operated hydraulic jack in increments of 4 kN (0.9 kip)
onto the steel column stub until failure (refer to Fig. 4). 1.18 209.50 42.19 57.81
Central deflections were recorded and cracks were marked, 0.6 57.00 11.48 46.33
incrementally, using two digital gauges positioned adjacent 0.3 21.50 4.33 42.00
to the column.
0.15 168.80 33.39 8.61
≤0.15 42.80 8.62 0.00

ACI Structural Journal/January 2018 165


Fig. 5—Flexural reinforcements.
Flexural cracks were observed on the tension surface of
the specimens at approximately 43% of the failure loads for
ISW M6, 50% of the failure loads for ISW M1 to M4, and
56% of the failure loads for ISW M5, as the eccentricity
decreased from 150 mm (5.9 in.) to 100 mm (3.9 in.) to
50 mm (2 in.), respectively. This is to be expected because
tensile stress would be increased as the column eccentric-
ities increased. Similar to flat slabs,1,20 as the applied load
continued to increase, torsional cracks on the tension surface
were observed to develop from the column side faces at an
angle of approximately 45 degrees.

Failure loads
The punching failure loads from tests, Ptest, are presented
in Table 3. It is important to note that the self-weight of the
Fig. 6—Typical load-deflection curves of Specimen IWS M2. slab specimens has not been included due to the small size
of slab specimens (limited to the points of contraflexure).
APPLICABILITY OF SMALL-SCALED SPECIMENS The shear capacity of a waffle slab was observed to increase
In view of the fact that the slab specimens described are as the size of the solid section increased. In comparison
1/10th-scaled specimens cast from micro-concrete having with Specimen ISW M3, cast with a 200 mm (7.9 in.) solid
a maximum aggregate size of 2 mm (0.08 in.), and rein- section, Specimens ISW M1 and M2 provided increments
forced with plain steel bars with small diameter 3.4 mm in punching strength of 18% and 23% as the solid section
(0.13 in.). A discussion with regards to the size effects on increased to 290 and 470 mm (11.4 and 18.5 in.), respec-
concrete strength in shear and in compression, the behavior tively. This is to be expected because of the increase in the
of micro-concrete in punching shear, and the bond strength revolution failure surface that has formed and mobilized
of small steel bars has been presented in an earlier paper.11 within the solid sections.
The shear capacity of the waffle slab(s) was observed to
TEST RESULTS decrease as the column eccentricity increased. In comparison
Failure modes with Specimen IWS M5, loaded with 50 mm (2 in.) eccen-
All slabs failed suddenly by punching shear in a mode tricity, Specimens IWS M4 and M6 exhibited reductions in
very similar to the mechanism observed in flat slabs.15-18 The punching strength of 17% and 42% as the applied eccentric-
failure surface was characterized by shear cracks inclined at ities increased to 100 and 150 mm (3.9 and 5.9 in.), respec-
approximately 22 degrees and intersected with the top slab at tively. It is believed that the observed reductions in shear
a distance of approximately 2.5 times the overall depth of slab capacities were attributed to the increase in the moment trans-
from the column face(s) comparable to that observed in flat ferred. In addition, when the applied moment exceeded the
slabs.19 For slabs loaded with shear span less than 2.5 times connection moment resistance, it transformed into torsion
the overall slab depth, the failure surface was observed to stresses on both sides of the column. As a result, nonuni-
propagate from the column faces to the supports as expected. formly distributed shear stresses occurred at the perimeter of
The shear cracks were limited to the front (heavily loaded) the column, thus leading to a lower punching shear capacity
region and two side regions of the column, while no cracks to be mobilized.
were observed in the rear region (refer to Fig. 7). However,
unlike a flat slab, when the width of solid section is less than Deflection
five times the total depth of the slab, the failure surface was The defection of each specimen was measured using two
an incomplete failure surface because some of the potential digital gauges placed at the front and the back faces of the
failure surface was lost due to the presence of waffle sections column after each load increment. A typical set of load-
(refer to Figure 7(I)).

166 ACI Structural Journal/January 2018


Fig. 7—Failure mechanism of waffle slab specimens.
Table 3—Comparisons between predicted and observed shear strengths
Slab Mtest*, kN.m (k.ft) Ptest†, kN (kip) Ppred, kN (kip) PACI, kN (kip) PEC2, kN (kip) PBS, kN (kip) PPred/Ptest PACI/Ptest PEC2/Ptest PBS/Ptest
IWSM1 4.21 (3.1) 42.1 (9.5) 33.5 (8) 27.5 (6.2) 46 (10.3) 58.0 (13) 0.89 0.65 1.09 1.38
IWSM2 4.42 (3.3) 44.2 (9.9) 47.0 (10.6) 30.5 (6.9) 48.6 (10.9) 61.3 (14.4) 1.06 0.69 1.10 1.39
IWSM3 3.58 (2.6) 35.8 (8.1) 36.8 (8.3) 34.7 (7.8) 51 (11.5) 63.9 (14.4) 1.03 0.97 1.42 1.79
IWSM4 4.21 (3.1) 42.1 (9.5) 39.5 (8.9) 32.3 (7.3) 49.7 (11.2) 62.6 (14.1) 0.94 0.77 1.18 1.49
IWSM5 2.53 (1.9) 50.5 (11.4) 44.7 (10) 38.3 (8.6) 53.1 (11.9) 66.2 (59.1) 0.92 0.76 1.05 1.31
IWSM6 4.42 (3.3) 29.47 (6.6) 37.2 (8.4) 28.2 (6.3) 46.8 (10.5) 59.1 (13.3) 1.26 0.96 1.59 2.00
Mean 1.02 0.80 1.24 1.47
Standard deviation 0.14 0.13 0.22 0.19
*
Mtest is applied moment.

Ptest is applied vertical load.

deflection curves recorded from the test of Specimen (IWS shear strength is obtained by equating the work done by the
M2) is presented in Fig. 6. Full deflection data are reported applied load to the energy dissipated in the failure surface.
in Reference 21. The load-deflection curve can be consid- However, due to its complexity, it is not possible to obtain a
ered as two straight lines inclined at two different angles.22 closed form solution. In view of this, Salim and Sebastian24
The first slope corresponds to the stiffness of an uncracked have developed a simplified approach based on a straight
section, while the second slope corresponds to the stiffness line failure surface that represents the complex method
of a cracked section.22 Such a phenomenon is consistent adequately but with a slightly greater scatter in the ratio of
with flat slabs failed by eccentric punching shear.15,17,18 It is test to predicted values. The simplified approach is adopted
believed that such a phenomenon is due to the fact that the in the current paper.
steel reinforcements become fully mobilized and effective The proposed model (refer to Fig. 8) considers that any
on nominal slip. change in slab thickness entered into the failure surface
incurs a reduction in the shear area and, consequently, a
PROPOSED MODEL reduction in the energy dissipation and hence the ultimate
General punching capacity. Thus, the reduction in punching capacity
The proposed model is based on the upper-bound plasticity depends on the loss of shear area, which in turn depends on
theory in predicting the punching shear strength of concrete the size of solid section. If the latter is wide enough to accom-
slabs. The failure surface of revolution is predicted to be a modate the punching revolution, then no reduction would be
catenary, possibly jointed by a straight line inclined at the required. The implication is that the model could predict the
concrete angle of friction to the plane of slab.23 The punching punching shear resistance of a flat slab by assuming either

ACI Structural Journal/January 2018 167


Fig. 8—Proposed eccentric punching shear failure surface.

Fig. 9—Eccentric punching failure surface of waffle slab.


the size of solid section is sufficiently large or the top slab is pated on the failure surface23 in MPa; and Ac is the punching
sufficiently thick. As shown in Fig. 8 and Appendix A,* the failure surface in mm.
punching shear strength of the waffle slab is computed based The details of calculation for shear failure surface areas and
on the following geometrical categories: summation of energy dissipated to determine the punching
1. Failure surface extends into the waffle section causing a capacities are given in Appendix A. The failure surface of
reduction in the shear capacity; revolution was found to be dependent on the column eccen-
2. Failure surface extends into the waffle section but not tricity e—that is, it was observed to be restricted to the front
resulting in loss in the shear capacity. This mechanism is (heavily loaded) region and the adjacent side regions of the
derived from the tests of Lau and Clark on ribbed slabs,25 column when the eccentricity was 1.0 and 1.5 times the
and shown herein for illustration as a possible failure mech- column size cx (refer to Fig. 9 and 10). When the eccentricity
anism; and was 0.5 times the column size cx, the failure surface was
3. The solid section is sufficiently wide, resulting in no observed to extend to the rear (lightly loaded) region of the
loss in the shear capacity. column (refer to Fig. 9(a)). As such, it can be concluded that
By categorizing the shear failure surface, the potential the cracks formation is inversely proportional to the eccen-
punching shear capacity, Ppot, can be computed from the tricities (refer to Fig. 11). These observed mechanisms are
following equation replicated in the proposed punching failure surface Ac in the
form of an effectiveness factor ξ that allows for the reduction
Pexδ = Pinδ = Ppot = αwi Ac (2) on effective shear area in the rear (lightly loaded) regions.
The effective punching shear failure surface Ac is expressed
where Pex is the external work done by the applied load in as follows
kN; δ is the relative displacement field between two rigid
zones; α is a shear retention factor; wi is internal work dissi-  π π 
Ac = h 1 + cot 2 θ  2cx + c y + h cot θ + ξ h cot θ (3)
 2 2 
*
The Appendix is available at www.concrete.org/publications in PDF format, where h is the total depth of slab; θ is the inclination angle
appended to the online version of the published paper. It is also available in hard copy
from ACI headquarters for a fee equal to the cost of reproduction plus handling at the of the failure surface; cx is the column size parallel to the
time of the request.

168 ACI Structural Journal/January 2018


Fig. 10—Internal shear cracks of Specimens IWS M4, M5, and M6.

Fig. 11—Schematic sketches of observed eccentric punching shear mechanism.


direction of the column eccentricity; cy is the column size sive strength to allow for the fact that concrete is not a
perpendicular to the column eccentricity; ξ is the effective perfectly rigid plastic material.26,27 Although the effective-
shear area factor ξ = 1– e/cx, e/cx ≤ 1; e is the column’s ness factor has been proposed to be a function of concrete
eccentricity, e = (Mtest/Ptest); Mtest is the transferred moment; compressive strength by Braestrup et al.,23 Salim and Sebas-
and Ptest is the vertical load. tian24 found that good correlation with the test results was
only achieved when Sigurdsson’s effectiveness factor28 was
Effectiveness factor used in their proposed models. As proposed earlier,11 the
In applying plastic theory to concrete, it is necessary to effectiveness factor is to be a function of concrete strength,
introduce an effectiveness factor vc to the concrete compres- flexural reinforcement ratio, and total depth of the slab

ACI Structural Journal/January 2018 169


of the 10 and 20 mm (0.4 and 0.8 in.) aggregates specimens
remained at their peak values, as indicated in Fig. 12. In view
of the fact that the tests described in the present paper were
carried out on micro-concrete slabs with a maximum aggre-
gate size of 2 mm (0.08 in.), a shear retention factor that
is dependent on the maximum aggregate size was adopted.
Values of 1.0 and 0.7 were adopted for normal concrete and
micro-concrete, respectively.

Moment transfer mechanism


In addition to the observed mechanism on the formation
of the shear failure surface, the torsion factor is another
important factor to be considered in predicting the eccen-
tric punching capacity. This is because when the applied
moment Mtest exceeds the flexural capacity Mf, the excessive
moment, (Mtest – Mf), is assumed to be redistributed within
the column side strips as torsion, T (refer to Fig. 13); subse-
quently, an additional force ΔP is induced at the inner face of
the column. The total downwards applied force is therefore
equal to (PPot + ΔP), and consequently, the punching failure
is assumed to take place when the total applied force is equal
to the punching shear resistance of the slab. The implication
is that when the moment transfer exceeds the slab’s flexural
Fig. 12—Shear stress-displacement curve.29,30 capacity, the punching failure would occur at a lower failure
load. This effect is accounted for in the proposed model
by the torsion factor β, which can be calculated using the
following equation

PPot
β= ; β ≤ 1 (6)
( PPot + ∆P )
where PPot is the potential punching shear strength to be
computed from Eq. (2), and ΔP is the additional downward
force induced by torsion.
The additional torsion shear stresses, τt, and the additional
induced downward force, ΔP, can be derived using the sand
heap analogy32 as

Fig. 13—Additional force ΔP induced by torsion stresses. 2T


τt = (7)
 h 
0.9  0.5  hmin 2  hmax − min 
vc =  0.7 +  (1 + 0.14ρ) (4)  3 
f cc h
Hence
where fcc is the cylindrical compressive strength of concrete
T
in MPa; h is the total depth of slab in meters; and ρ is the ∆P = if cx ≥ h (8)
average flexural reinforcement ratio. h
cx −
3
Shear retention factor α
In the current paper, a shear retention factor α was adopted  c 
T h − x 
from the shear tests of Boswell and Wong29 and Wong30 that  2
∆P = if cx < h (9)
were conducted on Mattock-type pushoff specimens31 cast  c 
cx  h − x 
with different maximum aggregate size of 2, 10, and 20 mm  3
(0.08, 0.4, and 0.8 in.). The results indicate that, although
the peak strength is essentially independent of the aggregate Finally, the eccentric punching shear capacity can be
size, the residual strength was dependent on the aggregate computed from the following equation.
size. With increasing shear displacements, the shear resis-
tance of 2 mm (0.08 in.) aggregate specimen reduces to PPred = βPpot (10)
approximately 70% of its peak value, while the resistance

170 ACI Structural Journal/January 2018


Flexural capacities Mf Table 4—Comparison between predicted eccentric
In deriving the flexural resistance, Mf, reference is made punching shear strength and test results of flat slabs
to the test observations of previous research,15,17,33,34 where PPred/Ptest
strains of the flexural reinforcements (parallel to the applied
Mean/standard deviation
moment) were measured to indicate the effective width of Total number
the slab in resisting the applied moment. It was noted that Researcher of slabs Proposed model
the tensile stress in the reinforcements decreased linearly up Ghali et al.21
3 1.01/0.17
to a distance equal to the effective depth, d, of the slab from Regan et al.22 16 0.92/0.10
both sides of the column. From these observations, the effec-
Hawkins et al.23 25 0.96/0.17
tive breadth (cy + d) is therefore proposed.33,35,36 In view of
the fact that the slab specimens tested in the current research Ozden et al. 24
6 0.83/0.08
consisted of moment transfer similar to those reported, and Total 50 0.94/0.15
hence, the proposed effective breadth (cy + d) is adopted.
The flexural resistance is based on the plasticity equa- slabs, as similar observations were made by others35-37 when
tions of pure bending derived for rectangular beams.27 The predictions derived from ACI8 were compared with flat slabs.
total flexural resistance of the connection, Mf, is the sum of The current EC29 overestimated the punching capac-
the flexural resistances developed at front and back of the ities of waffle slab specimens. This may be attributed to
column Mf1, Mf2 changes in waffle slab thickness and changes in the column
eccentricity. Comparisons with specimens having a small
Mf = Mf1 + Mf2 (11) solid section (ISW M3) and larger solid section (ISW M2)
revealed that EC29 overestimated the punching capacity (up
Comparisons with test results to 42%) of that having smaller solid section and overesti-
The proposing model develops the previous model11 to mated the punching capacity (up to 10%) of those having
predict the eccentric punching shear strength. The model a larger solid section. In comparison with Specimen IWS
was first used to predict the failure load of 40 flat slabs cast M5 (loaded with 50 mm column eccentricity), EC29 overes-
from normal concrete and subjected to eccentric punching timated the punching shear strength of Specimens IWS M4
shear.15-18 The mean ratio of the estimated strengths to tests and M6 by 18% and 59% as the eccentricities increased to
was 0.94 with a standard deviation of 0.15 (refer to Table 4). 100 and 150 mm (3.9 and 5.9 in.), respectively.
The confidence in adopting the shear retention factor of 0.7 Similar overestimations were observed from the use of
was derived from the recent works of Lau and Clark.25 Hence the current BS811010 in predicting the punching capaci-
the value of 0.7 was adopted in the model when comparing ties of waffle slabs. This may also be attributed to the two
with the current data. Good agreements were achieved aforementioned reasons. Comparisons between specimens
between the estimated strengths Ppred and the test results having a small solid section (ISW M3) and larger solid
(refer to Table 3). The mean ratio of predicted strength to section (ISW M2) revealed that BS811010 overestimated the
test was 1.02, with a standard deviation of 0.14. punching capacity (up to 79%) of those having a smaller
solid section and overestimated the punching capacity (up
EVALUATION OF CURRENT CODES APPROACH to 39%) of those having a larger solid section. Comparisons
The current codes of practice such as ACI 318-14,8 EC2,9 between specimens loaded with a small column eccentricity
and BS811010 have adopted the control surface approach for (IWS M5) and specimen loaded with larger column eccen-
design of flat slabs against punching shear mechanism. The tricity (IWS M6) revealed that BS811010 overestimated the
control failure surface is essentially a virtual vertical shear punching capacities from 31% to 100% as the eccentricities
surface at an assumed distance from the column faces. In the increased from 50 to 100 mm (2 to 3.9 in.), respectively.
presence of moment transfer, the shear stresses induced from
the vertical load combine with those from the transferring CONCLUSIONS
moment. The ultimate punching resistance is therefore the 1. The punching failure mechanism of a waffle slab was
sum of all shear stresses on the control surface. found to be similar in nature to that of a flat slab. However,
In Table 3, comparisons with test results were carried out for a slab with a width of solid section less than five times
to evaluate the applicability of design codes8-10 on eccen- the overall depth of slab, the observed shear failure surface
tric punching mechanism of waffle slabs. These predictions was incomplete due to the losses in the failure surface when
were calculated with the partial safety factor for the mate- it extended into the waffle section.
rial being set to unity and the shear retention factor α taken 2. In the presence of moment transfer, the punching failure
as 0.7. surface was observed to concentrate in the front heavily
In comparisons with both EC29 and BS8110,10 ACI8 loaded region and the side regions of the loading column.
provided conservative predictions toward the observed In addition, the punching shear capacity of waffle slabs was
punching failure loads on waffle slabs despite the fact that its observed to decrease as the column eccentricity increased.
critical section does not resemble the punching revolution, 3. The proposed upper-bound plastic model based on Niel-
and also does not consider the loss of failure surface on the sen-Braestrup’s shear analysis, coupled with the simplified
punching revolution. However, the conservativeness derived failure mechanism, has shown good agreement with the
from ACI8 is not limited to punching mechanism on waffle

ACI Structural Journal/January 2018 171


test results of flat slabs reported in the literature and the test Φ = internal friction angle of concrete = Φ ≅ 37 degrees
φ = mechanical degree of steel reinforcement, φ = (ρc/fcb)
results of the waffle slabs presented in the current work. λ = material properties, λ = 1 – (k – 1)ft/fc
4. Because the punching shear design procedure in the μ = material properties, μ = 1 – (k + 1)ft/fc
current code of practice does not consider the punching shear θ = angle of failure surface
ρ = average flexural reinforcement of solid section
mechanism of waffle slabs, comparisons with the test results ρc = flexural reinforcement ratio passing through effective bending
revealed that both the BS8110 and EC2 overestimated the width
punching failure loads, while ACI remained conservative. τt = torsion stress
υb = effectiveness factor of concrete in bending, MPa
ξ = effective shear area factor, ξ = 1– e/cx, e/cx ≤ 1
AUTHOR BIOS
Ahmed Faleh Al-Bayati is a Lecturer at Al-Nahrain University, Baghdad,
Iraq. He received his PhD from the University of Nottingham Malaysia REFERENCES
Campus, Semenyih, Malaysia. His research interests include the design 1. Regan, P. E., and Braestrup, M. W., “Punching Shear in Reinforced
of reinforced concrete structures and the punching shear strength of slab- Concrete: A State-of-the-Art Report,” Bulletin d’Information No. 168,
column connections. Comité Euro-International du Béton, Lausanne, Switzerland, 1985, 232 pp.
2. Magura, D. D., and Corley, W. G., “Tests to Destruction of a Multi
Lau Teck Leong is an Associate Professor at the University of Nottingham Panel Waffle Slab Structure 1964-1965 New York World’s Fair,” ACI
Malaysia Campus. He received his BEng and PhD from the University of Journal Proceedings, V. 69, No. 9, Sept. 1971, pp. 699-703.
Birmingham, Birmingham, UK. His research interests include the design 3. Xiang, X. Z., “Punching Shear Strength of Waffle Slabs at Internal
of reinforced concrete structures and the punching shear strength of slab- Columns,” PhD dissertation, University of Leeds, Leeds, UK, 1993.
column connections. 4. Hussein, A. F., “Punching Shear Strength at Edge Columns in Ribbed
Flat Slabs,” PhD dissertation, University of Leeds, Leeds, UK, 1994.
Leslie A. Clark is an Emeritus Professor of structural engineering at the 5. Mosalam, K. M., and Naito, C. J., “Seismic Evaluation of Grav-
University of Birmingham. His research interests include the design, anal- ity-Load-Designed Column-Grid System,” Journal of Structural
ysis, assessment, and behavior of concrete structures. Engineering, ASCE, V. 128, No. 2, 2002, pp. 160-168. doi: 10.1061/
(ASCE)0733-9445(2002)128:2(160)
ACKNOWLEDGMENTS 6. Benavent-Climent, A.; Cah’ıs, X.; and Catal’an, A., “Seismic
The authors would like to thank the University of Nottingham Malaysia Behaviour of Interior Connections in Existing Waffle-Flat-Plate Struc-
Campus for providing the financial support to this project. tures,” Engineering Structures, V. 30, No. 9, 2008, pp. 2510-2516. doi:
10.1016/j.engstruct.2008.02.004
7. Souza, S. S. M., and Oliveira, D. R. C., “Reinforced Concrete Waffle
NOTATION Flat Slabs under Shearing,” IBRACON Structures and Materials Journal,
Ac = assumed failure surface V. 4, No. 4, 2011, pp. 610-625.
As1&2 = top and bottom flexural reinforcement areas passing through 8. ACI Committee 318, “Building Code Requirements for Structural
effective bending width Concrete (ACI 318-08) and Commentary,” American Concrete Institute,
ax = ay shear spans in x- and y-axis Farmington Hills, MI, 2008, 473 pp.
b = effective bending width, b = cy + d 9. Eurocode 2, “Design of Concrete Structures—Part 1-1: General Rules
brib = rib width and Rules for Buildings,” CEN, EN 1992-1-1, Brussels, Belgium, 2004,
cx, cy = size of column parallel to x- and y-axis, respectively 225 pp.
Dx, Dy = sides of solid section parallel to x- and y-axis, respectively 10. British Standards Institution, “Structural Use of Concrete: Part 1,
d = effective depth of slab Code of Practice for Design and Construction, (BS 8110: Part 1: 1997),”
e = column’s eccentricity, e = (Mtest/Ptest) London, UK, 1997, 120 pp.
fc = effective concrete compressive strength in plasticity analysis, 11. Al-Bayati, A. F.; Lau, T. L.; and Clark, L. A., “Concentric Punching
fc = vcfcc Shear of Waffle Slab,” ACI Structural Journal, V. 112, No. 5, Sept.-Oct.
fcb = effective compressive strength in bending, fcb = υbfcc 2015, pp. 533-542. doi: 10.14359/51687906
fcc = cylinder compressive strength of concrete 12. Binici, B., and Bayrak, O., “Upgrading of Slab-Column Connections
fcu15 = concrete compressive strength measured from 15 mm cube Using Fiber Reinforced Polymers,” Engineering Structures, V. 27, No. 1,
fcu50 = concrete compressive strength measured from 50 mm cube 2008, pp. 97-107. doi: 10.1016/j.engstruct.2004.09.005
ft = effective concrete tension strength in plasticity analysis, ft/fc = 13. Johnson, R. P., “Strength Tests on Scaled Down Concrete Suitable
1/400 for Models with a Note on Mix Design,” Magazine of Concrete Research,
h1 = thickness of topping slab V. 14, No. 40, 1962, pp. 47-53. doi: 10.1680/macr.1962.14.40.47
h2 = total depth of slab 14. Endersbee, L. A., “Discussion of ‘Size Effects in Small-
k = concrete parameter, where Φ = 37 degrees Scale Models,’” ACI Journal Proceedings, V. 63, No. 11, Nov. 1966,
Mf = flexural capacity of connection Mf = Mf1 + Mf2 pp. 1572-1573.
Mf1 = flexural capacity at column front face, 15. Hawkins, N. M.; Bao, A.; and Yamazaki, J., “Moment Transfer
 1   1  from Concrete Slabs to Columns,” ACI Structural Journal, V. 86, No. 6,
2
M f 1 = 1 − ϕ As1 f y d1 = 1 − ϕ ϕb1d1 f cb Nov.-Dec. 1989, pp. 705-716.
 2   2 
16. Ghali, A.; Elmasri, M.; and Dilger, W., “Punching of Flat Plates
Mf2 = flexural capacity at column back face, under Static and Dynamic Horizontal Forces,” ACI Journal Proceedings,
 1   1  V. 73, No. 10, Oct. 1976, pp. 566-572.
M f 2 = 1 − ϕ As 2 f y d 2 = 1 − ϕ ϕb2 d 2 2 f cb 17. Regan, P. E.; Walker, P. R.; and Zakaria, K. A. A., “Tests of Rein-
 2   2 
forced Concrete Flat Slabs,” CIRIA Project RP 220, School of the Environ-
Mtest = transferred moment ment Polytechnic of Central London, London, UK, 1979, 217 pp.
PACI = predicted failure load by ACI 18. Ozden, S.; Ersoy, U.; and Ozturan, T., “Punching Shear Tests of
PBS = predicted failure load by BS8110 Normal-Strength and High-Strength Concrete Flat Plates,” Canadian
PEC2 = predicted failure load d by EC2 Journal of Civil Engineering, V. 33, No. 11, 2006, pp. 1389-1400. doi:
PPot = potential punching shear strength 10.1139/l06-089
PPred = predicted failure load 19. Regan, P. E., “Behaviour of Reinforced Concrete Flat Slabs,” CIRIA
Ptest = test failure load Report 89, Construction Industry Research and Information Association,
T = torsion moment London, UK, 1981, 89 pp.
vc = shear effectiveness factor 20. Alexander, S. D. B., and Simmond, S. H., “Shear-Moment Interac-
1 tion of Slab-Column Connections,” Canadian Journal of Civil Engineering,
wi = internal work dissipated wi = δf c ( λ − m cos θ ) V. 15, No. 5, 1988, pp. 828-833. doi: 10.1139/l88-108
α = shear retention factor 2
21. Al-Bayati, A. F., “Punching Shear Mechanisms of Waffle Slabs,”
β = torsion factor PhD dissertation, University of Nottingham Malaysia Campus, Malaysia,
ΔP = additional downward force induced by torsion 2013.
δ = displacement field in plasticity analysis

172 ACI Structural Journal/January 2018


22. Marzouk, H., and Hussein, A., “Experimental Investigation on the ings of Cement and Concrete Association Research Seminar, Wexham
Behavior of High-Strength Concrete Slabs,” ACI Structural Journal, V. 88, Springs, 1981.
No. 6, Nov.-Dec 1991, pp. 701-713. 30. Wong, S. S., “Collapse Behaviour of Micro-Concrete Box Girders
23. Braestrup, M. W.; Nielsen, M. P.; Jensen, B. C.; and Bach, F., Bridges,” PhD thesis, City University, London, UK, 1998.
“Axisymmetric Punching and Reinforced Concrete,” Report No. R-75, 31. Mattock, A. H.; Ibrahim, I. O.; and Hofbeck, J. A., “Shear Transfer in
Structural Research Laboratory, Technical University of Denmark, Lyngby, Reinforced Concrete,” ACI Journal Proceedings, V. 66, No. 2, Feb. 1969,
Denmark, 1976, 33 pp. pp. 119-128.
24. Salim, W., and Sebastian, W. M., “Plasticity Model for Predicting 32. Clark, L. A., Concrete Bridge Design to BS 5400, Construction Press,
Punching Shear Strengths of Reinforced Concrete Slabs,” ACI Structural London and New York, 1983, 186 pp.
Journal, V. 99, No. 6, Nov.-Dec. 2002, pp. 827-835. 33. Kanoh, Y., and Yoshizaki, S., “Strength of Slab-Column Connections
25. Lau, T. L., and Clark, L. A., “Shear Transfer between Ribbed Slab Transferring Shear and Moment,” ACI Journal Proceedings, V. 76, No. 3,
and Internal Column,” Magazine of Concrete Research, V. 59, No. 7, 2007, Mar. 1979, pp. 461-478.
pp. 115-128. doi: 10.1680/macr.2007.59.7.507 34. Yamazaki, J., and Hawkins, N.M., “Behavior of Concrete Plates
26. Nielsen, M. P.; Brsestrup, M. W.; Jensen, B. C.; and Bach, F., Jointed to Columns,” Proceedings Japan Society of Civil Engineers,
“Concrete Plasticity-Beam Shear-Shear in Joints-Punching Shear,” Special No. 292, pp. 117-130.
Publication, Danish Society for Structural Science and Engineering, Struc- 35. Park, H., and Choi, K., “Improved Strength Model for Interior Flat
tural Research Laboratory, Technical University of Denmark, Lyngby, Plate-Column Connections Subject to Unbalanced Moment,” Journal
Denmark, 1978, 129 pp. of Structural Engineering, ASCE, V. 132, No. 5, 2006, pp. 694-704. doi:
27. Nielsen, M. P., and Hoang, L. C., Limit Analysis and Concrete Plas- 10.1061/(ASCE)0733-9445(2006)132:5(694)
ticity, third edition, CRC Press, Boca Raton, FL, 2011, 799 pp. 36. Tian, Y.; Jirsa, J. O.; and Bayrak, O., “Strength Evaluation of Inte-
28. Sigurdsson, T. G., “Punching Shear by Eccentric Loading,” MSc rior Slab-Column Connections,” ACI Structural Journal, V. 105, No. 6,
dissertation, Technical University of Denmark, Lyngby, Denmark, 1991. Nov.-Dec. 2008, pp. 792-700.
(in Danish) 37. Broms, C. E., “Design Method for Imposed Rotations of Inte-
29. Boswell, L. F., and Wong, S. S., “The Shear Behaviour of rior Slab-Column Connections,” ACI Structural Journal, V. 106, No. 5,
Micro-Concrete and Its Inclusion in a Proposed Yield Criterion,” Proceed- Sept.-Oct. 2009, pp. 636-645.

ACI Structural Journal/January 2018 173


ACI STRUCTURAL JOURNAL GENERAL INFORMATION

ACI Research and Academic Opportunities

This article details some of the opportunities for researchers and professionals upon becoming a part of the ACI community. This article
will outline the possibilities available to members, such as attending The ACI Concrete Convention and Exposition, viewing past technical
presentations, access to a vast abstract library, and ACI’s Call for Papers. Up-to-date information concerning these and additional
opportunites can be found at ACI’s website, www.concrete.org.

THE ACI CONCRETE CONVENTION ACI Convention Schedule


AND EXPOSITION City Location Dates
ACI Conventions give attendees the opportunity to partic-
Salt Lake City, UT, Grand America &
ipate in the development of industry codes and standards, March 25-29, 2018
USA Little America
learn about the latest in concrete technology, network
with leading concrete professionals, and fulfill potential Rio All-Suites Hotel &
Las Vegas, NV, USA October 14-18, 2018
Casino
continuing education requirements.
ACI technical and educational sessions, which are held Quebec City
Quebec City, QC,
Convention Centre & March 24-28, 2019
during ACI Conventions, provide attendees with the latest Canada
Hilton Hotel
research, case studies, best practices, and opportunities to
earn Professional Development Hours and Continuing Duke Energy
Convention Center &
Education Units. ACI committees, whose meetings take Cincinnati, OH, USA
Hyatt Regency
October 20-24, 2019
place during the ACI Convention, develop the standards, Cincinnati
reports, and other documents needed to keep those in the
Rosemont, IL, USA Hyatt Regency O’Hare March 29-April 2, 2020
industry up to date with the latest technology. Committee
meetings are open to all registered convention attendees. Raleigh Convention
Raleigh, NC, USA Center & Raleigh October 25-29, 2020
The ACI Convention takes place twice a year—once in Marriott
the fall and once in the spring. ACI reserves rooms at local
Hilton & Marriott
hotels and offers a discounted rate to members. Networking Baltimore, MD, USA March 28-April 1, 2021
Baltimore
and other nontechnical events are coordinated through ACI
and take place at each convention. Hilton Atlanta
Atlanta, GA, USA October 17-21, 2021
Downtown

TECHNICAL PRESENTATIONS AND DOCUMENTS


descriptions of submission requirements and policies can
Access to a vast abstract library, online educational
be found at www.concrete.org. ACI’s website also contains
presentations, webinars, and ACI education documents are
a detailed list of the date(s), sponsor(s), and location(s) of
often free for members or offered at a discounted rate. New
events calling for papers.
presentations and documents are always being added.
Guidelines for submitting technical papers for review
to either the ACI Structural Journal or the ACI Materials
CALL FOR PAPERS
Journal can be found at ACI’s website.
ACI is accepting the submission of papers for conven-
tions, committees, chapters, and subsidiaries. Detailed

1132 ACI Structural Journal/September-October 2016

174 ACI Structural Journal/January 2018


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 115-S14

Modeling Time-Dependent Deformations: Application


for Reinforced Concrete Beams with Recycled
Concrete Aggregates
by Adam M. Knaack and Yahya C. Kurama
This paper describes the development and validation of a time- of the numerical model is presented to estimate the time-de-
dependent fiber-based (that is, layered) numerical model for the pendent behavior of a series of 18 cracked and uncracked
service-load deflection analysis of reinforced concrete structures. reinforced concrete beams with recycled concrete aggre-
Specifically, the model is applied to analyze the effect of recycled gates (RCA) as replacement for natural coarse aggregates
concrete aggregates (RCA) on the deflections of beams. Previous
(for example, crushed stone and gravel).
research has shown that increased deflections (rather than reduced
strength) may be a greater limitation for the use of RCA to replace
natural coarse aggregates (for example, gravel and crushed lime- RESEARCH SIGNIFICANCE
stone) for increased sustainability. Analysis and design tools are By recycling old concrete to replace natural aggregates,
needed to quantify this increase in deflections so that limits on the it may be possible to substantially reduce the need for new
use of recycled aggregates can be established. To aid in quanti- aggregates, thus helping with the preservation of forested
fying these increased deflections, a new time-dependent concrete areas and riverbeds. The use of RCA in the United States has
fiber including creep and shrinkage strains was developed in the been limited to non-structural applications such as roadway
open-source structural analysis program, OpenSees (Open System subbase, although the quality of RCA is usually much
for Earthquake Engineering Simulation). This paper describes the higher than is required in these applications and the material
validation of the model, including recent data from service-load is widely available in recycling plants across the country.
tests of slender cracked and uncracked RCA concrete beams.
Increased service-load deflections are one of the greatest
The  model was able to predict the time-dependent deflections
limitations for the use of RCA in structural applications
of reinforced concrete structural members under various load
scenarios; however, the initial (instantaneous) deflections were because of greater effects of RCA on the concrete stiffness,
generally underestimated because of underestimations in the extent creep, and shrinkage than on the concrete strength (Knaack
of cracking. The model was also able to predict the total strains and Kurama 2013a, 2015a). The analytical model described
and increased neutral axis depth over time as a result of creep and in this paper is able to quantify this effect. The open-source
shrinkage strains, except for the tension strains that were underes- application can especially serve as a tool to design struc-
timated because of the inability of the model to accurately predict tures that satisfy allowable service-load deflection limits not
the amount of cracking. It was found that the shrinkage strains only for RCA concrete but also for conventional reinforced
had a large effect on the time-dependent deflections of the beam concrete structures.
test specimens, which was not a finding available from the exper-
imental measurements. While the paper focuses on RCA concrete
BACKGROUND
applications, the numerical model is a general-purpose tool that
Deflections of reinforced concrete flexural members
can be used to analyze the time-dependent axial-flexural deforma-
tions of conventional reinforced concrete structures as well. can increase up to three to four times as a result of time-
dependent long-term load effects (Espion 1988). This
Keywords: creep; deflection; fiber modeling; recycled concrete aggregate; increase can be exacerbated when using RCA concrete. For
reinforced concrete; service load; shrinkage. example, considerable increases have been observed in the
immediate and time-dependent deflections of RCA concrete
INTRODUCTION beam test specimens (Maruyama et al. 2004; Sato et al.
Previous research on the numerical modeling of time- 2007; Li 2009; Lapko and Grygo 2010; Malešev et al. 2010;
dependent deformations of concrete structures is extensive Ajdukiewicz and Kliszczewicz 2011; Choi and Yun 2013).
and has included both fiber element (that is, layered) as well The results from these studies are generally limited to a few
as continuum finite element approaches (for example, Lou tests using RCA and only one load level in each research
et al. [2016], Mercan et al. [2013], Bacinskas et al. [2012], program. For example, for a service-load duration of approx-
Sousa et al. [2012], Marí et al. [2010], Mazzotti and Savoia imately 15 weeks, Lapko and Grygo (2010) showed 20%
[2009], Chong et al. [2008], Maekawa et al. [2006], and greater deflections in RCA concrete beams with full (100%)
Scanlon and Murray [1974]). Different from these previous
ACI Structural Journal, V. 115, No. 1, January 2018.
research efforts, the current paper describes a new time-de- MS No. S-2017-019.R1, doi: 10.14359/51701153, was received January 29, 2017,
pendent fiber-based numerical model within the framework and reviewed under Institute publication policies. Copyright © 2018, American
Concrete Institute. All rights reserved, including the making of copies unless
of an open-source, general-purpose structural analysis soft- permission is obtained from the copyright proprietors. Pertinent discussion including
author’s closure, if any, will be published ten months from this journal’s date if the
ware (OpenSees 2006). Additionally, a specific application discussion is received within four months of the paper’s print publication.

ACI Structural Journal/January 2018 175


aggregate replacement as compared to natural aggregate is 1 day and the applied load is equal to the full load, P. The
(NA) concrete beams. These tests were conducted on small- existing pseudo-time concept in the OpenSees Static Analysis
scale beams with a cross section of 3 x 5 in. (76 x 127 mm). was advantageous for the development of the new time-
Ajdukiewicz and Kliszczewicz (2011) found similar results dependent material model in the sense that time was already
for beams with full aggregate replacement loaded for up to defined in the program for use as the time domain in the new
1 year. Choi and Yun (2013) found the ratios of long-term material model.
to immediate deflections for RCA concrete beams to be To allow for the inclusion of concrete creep and shrinkage
smaller than the ratios for conventional natural aggregate deformations during a pseudo-time step, a new analysis type
(NA) concrete beams. A more recent study involving 18 called Creep Analysis was implemented based on the Static
RCA concrete beams was conducted by Knaack and Kurama Analysis function already defined within the OpenSees
(2013b; 2015b). framework, and a new uniaxial concrete material type called
None of the aforementioned experimental studies included TDConcrete was developed based on the existing Concrete02
numerical modeling of the time-dependent deflections of material. Full information on the new capabilities imple-
RCA concrete structures. Various methods are available mented in OpenSees (including source codes, sample input
for the prediction of the long-term service-load deflections and analysis files, and element recorders) can be found in the
of reinforced concrete structures, including simple design final project report (Knaack and Kurama 2013b).
code-based methods, closed-form analytical solutions, and
detailed fiber element and continuum finite element models. Overview
Code approaches such as in ACI 318-14 and CEN (2004) The new time-dependent TDConcrete material in
are generally simple, but not accurate (Ghali 1993; Gilbert OpenSees uses the following assumptions:
1999). Many analytical and numerical approaches (Teng 1. Concrete is fully hardened and able to resist load 2 days
and Branson 1993; Nie and Cai 2000; Kaklauskas 2004; after casting.
Rodriguez-Gutierrez and Aristizabal-Ochoa 2007; Balev- 2. Because the analysis is carried out under service loads,
ičius 2010) have been demonstrated to provide consid- concrete is assumed to remain in the linear elastic range in
erable improvements on the deflection predictions. One compression.
important limitation of the available numerical models is 3. Concrete fibers exceeding the cracking strain are expected
that they have used commercial finite element programs (for to have some residual stresses through tension stiffening.
example, ABAQUS and DIANA) or specific in-house codes 4. As an important limitation, the strength and stiffness of
developed by the researchers. Different from these concrete are assumed to remain constant during the entire
approaches, the model in this paper uses simplifying assump- analysis (that is, increases in the concrete strength and stiff-
tions and approximations within a widely and freely available ness with age are not included). This is a reasonable approx-
structural analysis platform, OpenSees (2006). Furthermore, imation for structures loaded at or after 7 days of curing, as
to the best of the authors’ knowledge, this is the first detailed investigated in this paper and as would be most common in
numerical modeling study for RCA concrete beams. practice (due to presence of shoring). For structures loaded
prior to 7 days, the strength and stiffness gain of concrete can
TIME-DEPENDENT CONCRETE MATERIAL MODEL be significant (Neville 2011), and thus, numerical models
To result in a widely available and widely applicable that incorporate the effects of aging may be needed for more
analysis tool, a new concrete material model including accurate analysis.
time-dependent creep and shrinkage strains under sustained 5. There are no thermal strains (that is, no significant
service loads was implemented by adapting the existing temperature changes) during the analysis.
Static Analysis and fiber element modeling framework in the 6. The stress in each fiber remains constant during each
open-source structural analysis program, OpenSees (2006). pseudo-time step. As described later, this is a common
One of the advantages of fiber element models is the ability assumption that has been used in previous research (Zien-
to represent the axial-flexural behavior of concrete structures kiewicz et al. 1968; Kabir 1977; Kawano and Warner 1996).
using cross sections that are discretized into concrete and 7. The concrete creep strain in tension has an equal
steel fibers together with a uniaxial stress-strain relationship magnitude to the creep strain in compression for the same
for each fiber. More information on fiber element modeling magnitude of stress. While this is also a common assump-
for structural analysis can be found in Taucer et al. (1991) tion that has been used in the past (Kabir 1977; Gilbert and
and Jiang and Kurama (2010). Ranzi 2011), according to Neville (2011), creep of concrete
All of the available material models in OpenSees are in tension is generally greater than in compression, with
time-invariant, and as such, there is no built-in capa- differences as high as 100% for concrete stored at a relative
bility for a time-dependent creep and shrinkage analysis. humidity of 50% and loaded at early ages. As such, some of
However, the magnitudes of the applied loads during a Static the model discrepancies discussed later in this paper could
Analysis are defined using a “pseudo-time” concept (that have developed from an underestimation of the creep strains
is, time associated with the analysis domain). For example, in tension. However, Neville (2011) also states that reli-
consider a linear time series with a pseudo-time step of able conclusions about the relative magnitudes of creep in
0.1 days for an applied force, P reached at 1 day. After one tension and in compression cannot be made, citing contra-
analysis step, the pseudo-time in the model is 0.1 days and the dictory evidence that shows greater compression creep than
applied load is 0.1P. After 10 analysis steps, the pseudo-time tension creep.

176 ACI Structural Journal/January 2018


creep deformations. The last Creep Analysis for this loading
duration should be conducted at t = 14 days.
4. Static Analysis at t = 14 days to apply additional loads
from the slab above.
5. A series of Creep Analyses starting at t = 14 days to
accumulate creep and shrinkage deformations from t = 14
days to the end of the loading duration.
It can be seen that a new series of Static Analysis and
subsequent Creep Analyses is conducted for each increment
of applied loads (for example, if additional dead loads or
live loads are applied to the column at later times). Typically,
the best practice for each series of Creep Analyses is to use
pseudo-time steps (by explicitly specifying the pseudo-time
for each analysis using the OpenSees “setTime” command)
that are uniformly spaced in logarithmic scale (Bažant and
Wu 1973) so that more analyses are conducted when the load
is first applied (that is, when the creep strains increase at a
faster rate). The analyses can then be run using the “analyze”
command.
The “analyze” command in OpenSees will run one Creep
Fig. 1—Mechanical stress-strain behavior of concrete Analysis at the pseudo-time set by the user. To run a series
(compression negative). (Note: 1 MPa = 0.145 ksi.) of Creep Analyses in succession at additional time steps (to
continue accumulating creep and shrinkage deformations for
8. Once the concrete cracking strain has been exceeded,
the remainder of the loading duration), the “setTime” and
creep and shrinkage still occur according to the same consti-
“analyze” commands are put inside a loop in the OpenSees
tutive relationships as an uncracked fiber, but under the
input file so that the pseudo-time is set prior to each analysis
reduced post-cracking (that is, tension stiffening) stresses.
for the desired number of analyses. An example time-
To analyze the time-dependent behavior of a concrete
dependent analysis loop using logarithmically spaced time
structure, a series of Static and Creep Analyses are
steps is provided in Appendix D of Knaack and Kurama
conducted at discrete times throughout the loading duration.
(2013b).
For example, consider the following scenario for the loading
Note that when conducting a new series of Static
and construction of a reinforced concrete column:
Analysis and subsequent Creep Analyses, the entire load
1. t = 0 days: Start time of the analysis. Column is cast and
increment for that series must be applied instantaneously
moist cured for 7 days.
during the Static Analysis and then held constant for the
2. t = 2 days: Column is assumed to be fully hardened and
duration of the subsequent Creep Analyses. This requires
stresses are induced due to self-weight. Creep strains begin
the use of a constant time-series to define each applied
to accumulate.
load increment in OpenSees. If additional load increments
3. t = 7 days: Moist curing of column ends and shrinkage
are placed on the structure at a later time, a new series of
strains begin to accumulate.
Static Analysis and subsequent Creep Analyses should be
4. t = 14 days: Additional loads applied from the slab above.
conducted using a similar process.
An OpenSees analysis of this column can be conducted
The TDConcrete material defines the total concrete strain
by modeling the concrete using the TDConcrete material.
(excluding thermal strains) at time, t as
To conduct a time-dependent analysis, the user must set the
pseudo-time of the model so that a combination of Static
Analyses and Creep Analyses can be conducted at the desired total (t ) = �  m (t ) +  cr (t ) +  sh (t ) (1)
times to determine the initial and long-term deflections of
the column. To set the pseudo-time, the OpenSees “setTime” where ϵtotal is total strain; ϵm is mechanical strain; ϵcr is creep
command is used. For the aforementioned column construc- strain; and ϵsh is drying shrinkage strain.
tion scenario, the analysis would progress as follows: During a Creep Analysis step at time t, the strain compo-
1. Static Analysis at pseudo-time, t = 2 days to apply nents for each TDConcrete fiber are determined by
column self-weight. Because the column is cast at t = 0 days carrying out iterations until a specified level of conver-
and assumed to take load once fully hardened at t = 2 days, gence is achieved on the global unbalanced force vector
the pseudo-time at the beginning of the analysis is set to in OpenSees. The fiber shrinkage strain, ϵsh(t) is calcu-
2 days. lated using the difference between the time t for the current
2. A series of Creep Analyses starting at t = 2 days to accumu- analysis step and the time td at the start of concrete drying,
late creep deformations from t = 2 days to 7 days. Note that no and remains constant during all iterations within the time
shrinkage strains would accumulate during moist curing. step. The creep strain, ϵcr(t) for the current analysis step is
3. A new series of Creep Analyses starting t = 7 days to calculated using the stress from the previous analysis step
start accumulating shrinkage deformations in addition to and also assumed to remain constant for the time step. Then,
the mechanical strain is calculated by subtracting ϵsh(t) and

ACI Structural Journal/January 2018 177


Fig. 2—Concrete strains: (a) strain components (excluding thermal strain) under constant stress;
and (b) superposition of
creep strains under time-varying stress.
ϵcr(t) from the total strain, ϵtotal(t), and the concrete mechanical on the tension-stiffening envelope towards the origin. Any
stress-strain constitutive relationship is used to determine the subsequent reloading in tension also occurs along this secant
stress, σ(t) in the fiber. Finally, the stresses of all fibers within stiffness until a new maximum tension strain is reached.
the cross section are integrated and the internal forces within
the element are passed on to the global analysis to check Creep strain constitutive relationship
convergence of the structure unbalanced force vector. Figure 2(a) shows the three components of the concrete
Provided that the Creep Analysis time steps are sufficiently total strain, given initiation of drying at t = td and a constant
small so that there is only a small amount of stress change stress applied at t = t0. According to Gilbert and Ranzi
within each time interval, the aforementioned method leads (2011), creep strain is approximately proportional to stress
to a reasonably accurate determination of the creep strains. when the sustained stress is within the linear elastic limit.
A possibly more accurate alternative to this approach would When the stress is assumed to remain constant with respect
be to assume a linearly varying concrete stress within each to time (that is, dσ/dt = 0), the creep strain at time t can be
time step; however, this approach was not used in this expressed as
research because it was deemed unjustified given the many
inherent uncertainties on the assumed material properties  cr (t ) =  m (t ) ⋅ φ0 (t ) (2)
(for example, concrete strength and stiffness).

Mechanical stress-strain behavior  m (t ) = σ (t ) / Ec (3)


Figure 1 shows the assumed concrete mechanical stress-
strain behavior. As stated previously, the model uses a where ϕ0(t) is the creep coefficient at time t.
linear-elastic stress-strain relationship in compression. The Then the load-induced strain (that is, sum of mechanical
concrete stiffness, Ec and tension strength, ft′ are input by and creep strains) can be written as
the user and should typically be specified at the concrete age
when the superimposed service loads are first applied. The  m (t ) +  cr (t ) =  m (t ) ⋅ 1 + φ0 (t )  (4)
tension-stiffening behavior after cracking uses a model from
Tamai et al. (1988). A value for the tension-stiffening expo- As shown in Fig. 2(b) and explained in Gilbert and Ranzi
nent, bts is recommended by Tamai et al. (1988) as bts = 0.4 (2011), the creep strain at time t as a result of time-varying
for concrete with deformed reinforcing bars. This value is stress can be determined using the principle of superposition as
based on a study with steel reinforcement in the post-yield
range, and users can specify other values of bts for the linear
 φ0 (t ) ⋅ ∆σ 0 / Ec , t0 ≤ t < t1
steel material range if experimental data is available. Once 
the concrete is cracked, any load reversal is assumed to  φ0 (t ) ⋅ ∆σ 0 + φ1 (t ) ⋅ ∆σ1  / Ec , t1 ≤ t < t2
cr (t ) =   (5)
follow the secant stiffness from the maximum strain reached  
 φ0 (t ) ⋅ ∆σ 0 + …+ φ n (t ) ⋅ ∆σ n  / Ec , tn ≤ t

178 ACI Structural Journal/January 2018


Fig. 3—Validation under constant uniaxial stress: (a) model; and (b) comparisons for creep frame CS-11 from Knaack and
Kurama (2013b, 2015a). (Note: 1 mm = 0.0394 in.)
This relationship can also be written as Shrinkage strain constitutive relationship
Similar to the creep constitutive relationship, TDConcrete

n
cr (t ) = ∑φi (t ) ⋅ ∆σ i / Ec (6) uses ACI 209R-92 drying shrinkage model as
i =0

Gilbert and Ranzi (2011) found that the principle of super-  sh (t ) = (t − td ) / ( ψ sh + t − td ) ⋅  sh ,u (8)
position applied to creep strains agrees well with available
test data for increasing stress histories. However, for concrete where td is time at initiation of concrete drying; ψsh is
under decreasing stresses (that is, unloading), superposition shrinkage time-function parameter; and ϵsh,u is ultimate
tends to overestimate the amount of creep recovery, thus shrinkage strain. All ACI 209R-92 adjustment factors for
leading to reduced creep deformations. Regardless, the prin- non-standard conditions can be applied by modifying the
ciple of superposition provides a reasonable approximation ultimate shrinkage strain ϵsh,u provided as input.
of the creep strains for most practical purposes. The process
is assumed to be valid for both uncracked and cracked VALIDATION OF TIME-DEPENDENT
concrete. Note that in a real reinforced concrete structure, CONCRETE MODEL
the concrete immediately adjacent to a crack unloads and is Constant stress under uniaxial load
not capable of carrying any stress, which is not captured by The first validation study for the TDConcrete analysis
the tension stiffening modeling approach. package implemented in OpenSees was based on the uniaxial
Note also that the summation in Eq. (6) requires the entire RCA concrete creep test data from Knaack and Kurama
stress history of each fiber to be stored during the anal- (2013b, 2015a). In each creep test frame, three pairs of plain
ysis for each integration point along each element. While concrete cylinders were stacked and a constant axial load
not implemented in TDConcrete, the use of a Dirichlet was applied at a specific age. The three pairs of cylinders
series approximation (Bažant and Wu 1973; de Borst and represented concrete mixtures with three different aggre-
van den Boogaard 1994; Kabir 1977; Taylor et al. 1970) gate volumetric replacement ratios (that is, R = 0%, 50%,
to the creep strain superposition could reduce computing and 100%; where R = 0% means no RCA in the concrete
times by allowing the concrete state (that is, stress and total mixture and R = 100% means no natural coarse aggregate
strain) to be determined from the previous time step rather in the mixture), and the results from each pair were aver-
than requiring integration at each time step over the entire aged to determine the total strain over the loading duration.
strain history. The analytical model for this experimental configuration is
The creep coefficient at time t is assumed to follow shown in Fig. 3(a).
ACI 209R-92 model as During each test, shrinkage strains were simultane-
ously measured using companion cylinder pairs under no
applied load, and the load-induced strain (that is, creep plus
φ i ( t ) = ( t − ti ) /  ψ cr , 2 + (t − ti )  ⋅ φu (7)
ψ cr ,1 ψ cr ,1

  mechanical strain) of the creep cylinders was determined by
subtracting the shrinkage strain from the total strain. The
where ti is time at stress change; ψcr,1 and ψcr,2 are creep
mechanical strain was taken as the first data point immedi-
time-function parameters; and ϕu is ultimate creep coeffi-
ately after load application, and the creep strain was deter-
cient. The creep time-function parameters ψcr,1 and ψcr,2 and
mined as the load-induced strain minus this mechanical
the ultimate creep coefficient ϕu are specified by the user as
strain. The creep coefficient was calculated as the creep strain
inputs to TDConcrete. Because the creep model in Eq. (7)
divided by the mechanical strain and a nonlinear regression
is for a standard condition corresponding to moist-cured
analysis was conducted using Eq. (7) to provide the best fit
concrete loaded at an age of 7 days, all of the modifications
to the data. The resulting ψcr,1, ψcr,2, and ϕu parameters were
(multipliers) specified in ACI 209R for non-standard condi-
then provided as input to TDConcrete. The concrete stiff-
tions can be applied by adjusting ϕu.

ACI Structural Journal/January 2018 179


Fig. 4—Validation under time-varying stresses: (a) model; and (b) comparison with Kawano and Warner (1996). (Note: 1 mm =
0.0394 in.; 1 m = 39.4 in.; 1 MPa = 0.145 ksi.)
ness  Ec was calculated from the mechanical strain. In the (1996) and TDConcrete for the varying concrete stresses or
analysis, the load was applied at the same age and magnitude the creep strains during the analysis.
as the test and was held constant over the entire duration.
As illustrative results, the analytical and experimental Time-varying stress under flexural load
load-induced strains for creep frame CS-11 (Knaack and Validation of TDConcrete under flexural load was
Kurama 2013b, 2015a) are shown in Fig. 3(b). The results conducted using the analytical example from Kabir (1977),
from the other creep frames exhibited similar comparisons. which was based on a beam specimen tested by Washa and
It can be seen that the TDConcrete model provided an excel- Fluck (1952). The beam was analyzed using the model in
lent fit to the test data by producing the creep strain model Fig. 5(a). The concrete compression strength and modulus of
that was used as input. This validates that for a constant elasticity (at an age of 14 days, when superimposed loading
stress, TDConcrete was able to correctly use the intended was applied), as well as the steel yield strength, were taken
creep formulation within the Creep Analysis implemented from measured values. Cracking was observed under the
in OpenSees. initial application of load; and thus, the concrete tension
strength was determined so that the initial deflection from
Time-varying stress under uniaxial load the analysis matched the measured deflection. A tension
The validation of TDConcrete under variable uniaxial stiffening exponent of bts = 0.4 was used as recommended
stresses involved the analysis of a reinforced concrete by Tamai et al. (1988). As shown in Fig. 5(b), the concrete
column previously analyzed by Kawano and Warner (1996). creep parameters, ψcr,1, ψcr,2, and ϕu were determined from
In this analysis, the column was subjected to a constant nonlinear regression analysis on the creep coefficient from
uniaxial load, but time-varying stresses were present as the concrete cylinders tested by Washa and Fluck (1952).
redistribution of stresses between concrete and reinforcing Similar to Kabir (1977), shrinkage strain was neglected in
steel (reduction of concrete stresses and increase of steel the analysis.
stresses) occurred due to concrete creep and shrinkage. As The beam deflection results depicted in Fig. 5(c)
shown in Fig. 4(a), the analytical model consisted of a single demonstrate an excellent fit with Kabir (1977)
element with one TDConcrete fiber and eight steel fibers over the entire time history. Both the TDConcrete and Kabir
(using the Steel01 material in OpenSees) to represent the analyses gave conservative (that is, greater) beam deflec-
column reinforcement. Figure 4(b) shows no differences tions in comparison with the test data from Washa and Fluck
between the analysis predictions from Kawano and Warner (1952) for approximately the first 100 days of loading,
which could be because the effects of aging on the concrete

180 ACI Structural Journal/January 2018


Fig. 5—Beam under flexural load: (a) model; (b) regression fit to creep data; and (c) comparisons for midspan deflection.
(Note: 1 mm = 0.0394 in.; 1 m = 39.4 in.)
strength and stiffness were ignored. After this loading were subjected to service-level loads, the steel stresses
period, both analyses provided reasonable estimates of the remained in the elastic range. The concrete in compression
measured data. These results support the use of TDConcrete also remained within the elastic range (that is, σ < 0.50fc′).
in the flexural analysis of reinforced concrete structures. The tension-stiffening exponent was chosen as bts = 0.8 to
result in the best match to the initial and time-dependent
VALIDATION AND APPLICATION FOR BEAMS deflections. Note that the use of a greater bts value for these
WITH RCA CONCRETE specimens as compared with bts = 0.4 for the specimen in
This section describes the application of TDConcrete to the Fig. 5 indicates reduced tension-stiffening stresses in the
RCA concrete beams tested by Knaack and Kurama (2013b, concrete (refer to the tension-stiffening equation in Fig. 1).
2015b). As shown in Fig. 6(a), each l = 4.0 m long (13 ft) This may be expected because the smaller total area and
beam was subjected to four-point bending using a concrete greater diameter of the two M16 bars for the specimens
weight block to simulate superimposed service loads for a in Fig. 6 likely resulted in smaller tension stress transfer
period of at least 119 days. The beams were b = 150 mm (through bond) to the concrete than the four M13 bars for
(6.0 in.) wide and h = 230 mm (9.0 in.) deep. In the specimen the specimens in Fig. 5.
nomenclature (UT, UC, and CC series beams), the first letter Companion concrete cylinder creep and shrinkage test data
“U” indicates beams that were subjected to a superimposed for each beam, as described in Knaack and Kurama (2013b,
load designed not to induce immediate cracking, whereas 2015a), were used to determine the creep and shrinkage
the first letter “C” indicates beams that were subjected to input (using nonlinear regression) into the TDConcrete
a greater superimposed load to induce immediate cracking. model. Six beams (UT-0-28, UT-0-7, UT-50-28, CC-50-7,
For the second letter of the nomenclature, the letter “T” indi- CC-100-28, and UC-100-7) did not have any companion
cates beams with tension (bottom) longitudinal bars only and creep and shrinkage measurements; therefore, the measured
“C” indicates beams with compression (top) longitudinal creep and shrinkage data for other beams were used as input
bars and transverse stirrups in addition to tension (bottom) for these specific beams. Although ACI 209-92 suggests the
longitudinal bars. Subsequent numbers in the nomenclature use of a size adjustment factor for creep and shrinkage, Euro-
indicate specimens with different amounts of coarse aggre- code 2 (CEN 2004) does not require a size adjustment factor
gate replacement (R = 100, 50, and 0% replacement) and for beams of this size. Therefore, it was decided to neglect
beams loaded at the age of t0 = 7 days versus 28 days (for any size adjustment factor in the creep and shrinkage models
example, UT-100-28 for a UT series beam with 100% aggre- for the beams. Also, because the creep and shrinkage cylin-
gate replacement and loaded at 28 days). ders were kept in the same room conditions and cured the
The average measured (on the day of superimposed same way as the corresponding beam specimens, no adjust-
loading) material properties fc′, ft′, and Ec, were used as input ment factor for humidity or curing needed to be applied.
to the TDConcrete analyses. Considering that the beams

ACI Structural Journal/January 2018 181


Fig. 6—Service-load deflection modeling of RCA concrete beams: (a) test elevation; (b) analytical model; and (c) loading
schedule. (Note: 1 mm = 0.0394 in.; 1 m = 39.4 in.)
As shown in Fig. 6(b), each beam was modeled in Figure 6(c) shows the loading sequence for the test beams.
OpenSees using a total of 20 displacement-based beam- From t = td = 2 days until t = t0 (that is, first application of
column elements with Gauss-Lobatto integration scheme superimposed service load), 20 pseudo-time steps uniformly
and three integration points along each element. The spaced in logarithmic scale were used to run a Static
Legendre integration scheme, also available in OpenSees, Analysis and subsequent series of Creep Analyses to deter-
provided the same results as the Gauss-Lobatto scheme. The mine the immediate and creep deflections under self-weight
beam cross-sections were discretized into 20 concrete fibers as well as the shrinkage deflections during this period.
and up to two (top and bottom) steel fibers depending on As discussed in Bažant and Wu (1973), the use of a loga-
the specimen. The load was applied in two steps: 1) uniform rithmic scale allows more time steps when the load is first
beam self-weight assumed to take effect at t = 2 days after applied and creep and shrinkage strains are increasing at a
casting (in the laboratory, each beam was placed on wooden faster rate. From t = t0 until the age at which the service
end blocks immediately after removal from formwork at load was removed at t = tr, another Static Analysis and Creep
2 days); and 2) two superimposed point loads applied at Analysis series with 50 pseudo-time steps uniformly spaced
±190 mm (±7.5 in.) from the midspan at a service loading in logarithmic scale were used to determine the additional
age of t0 = 7 days or 28 days, depending on the beam. It was immediate, creep, and shrinkage deflections during this
assumed that the drying shrinkage of each beam began at td = stage. Finally, from t = tr until the end of measurements
2 days (age at which the beams were removed from the form- at t = te, the number of time steps was varied based on the
work) and that the shrinkage strains were uniformly distrib- duration. A few beams had an additional superimposed load
uted throughout the entire beam volume (that is, all of the applied during the last 2 months of testing (dashed lines in
concrete fibers were given the same shrinkage properties). Fig. 6(c)). This stage of increased loading was modeled using
30 logarithmically spaced time steps. Preliminary analyses

182 ACI Structural Journal/January 2018


Fig. 7—Deflection comparisons for UT series beams: (a) UT-0-7; (b) UT-0-28;
(c) UT-50-7; (d) UT-50-28; (e) UT-100-7; and
(f) UT-100-28. (Note: 1 mm = 0.0394 in.)
were conducted to ensure that the selected number of beam- not specify an adjustment factor for loading ages less than
column elements, fibers, and time steps were adequate for 7 days, Eq. (10) was also used for loading ages less than Tcr
the accuracy of the analyses. (to determine the creep strains under self-weight), resulting
The ACI 209R-92 loading age adjustment factor, with in adjustment factors greater than 1.0 for those conditions.
additional modifications (Knaack and Kurama 2013b) to The results from the time-dependent analyses are shown in
better represent the specific loading schemes from the exper- Fig. 7, 8, and 9 for the UT, UC, and CC series beams, respec-
imental program, were used in the analyses as follows. tively. Similarly, Table 1 compares the predictions from
For loading ages later than 7 days for moist cured concrete, TDConcrete with the measured deflections at t = t0 (imme-
ACI 209R specifies a loading age adjustment factor as diate deflection from the application of superimposed load
at t0) and at t = t0 +119 days (total accumulated deflection
γ la , ACI 209 = 1.25 ⋅ (ti − tcast ) from t0 until t0 +119 days), which were zeroed immediately
−0.118
(9)
before the superimposed loading at time, t = t0. Note that
where ti is time at stress change; and tcast is time at concrete measured deflection data for beams CC-50-28 and UT-100-7
casting. The ACI 209R age adjustment factor takes a value were not available due to failure of the displacement trans-
less than 1.0 for ti – tcast greater than 7 days, and is intended ducers; and thus, only the analytical predictions are shown
to be applied as a multiplier to the standard creep relation- for these beams. Note also that the analytical data in Fig. 7 to
ship in Eq. (7). For the purposes of the analyses herein, the 9 start at t = td = 2 days, whereas the measured data start at t
ACI 209R adjustment factor in Eq. (9) was adapted to a = t0. Because the laboratory measurements did not start until
more general relationship applicable to other loading ages as the application of the superimposed load at t = t0, the beam
deflections from t = td = 2 days to t = t0 (due to shrinkage and
the immediate and creep deflections due to self-weight) were
γ la = γ la , ACI 209 / 1.25 ⋅ Tcr−0.118  (10)
not measured. However, to accurately simulate the evolution
of concrete and steel stresses analytically, it was necessary
where Tcr is the loading age of the cylinders used to develop
to run the analyses over the entire loading history beginning
the creep relationship (that is, 7 or 28 days in this research).
at t = td = 2 days. The resulting analytical deflections were
In this more general relationship, γla = γla,ACI209 when Tcr =
subsequently zeroed at the application of the superimposed
7 days (that is, Eq. (10) is reduced to Eq. (9) when the creep
load at t = t0 to permit direct comparisons with the measured
relationship is based on specimens loaded at an age of 7 days),
data starting at t = t0.
and γla = 1 when ti – tcast = Tcr (that is, Eq. (10) gives a creep
The comparisons show that the analytical models were
adjustment factor of 1.0 when the age of concrete at stress
able to predict the time-dependent service-load deflections
change is the same as the loading age of the cylinders used
of the test beams reasonably well; however, the immediate
to develop the creep relationship). Because ACI 209R does

ACI Structural Journal/January 2018 183


Fig. 8—Deflection comparisons for UC series beams: (a) UC-0-7; (b) UC-0-28;
(c) UC-50-7; (d) UC-50-28; (e) UC-100-7;
and (f) UC-100-28. (Note: 1 mm = 0.0394 in.)

Fig. 9—Deflection comparisons for CC series beams: (a) CC-0-7; (b) CC-0-28;
(c) CC-50-7; (d) CC-50-28; (e) CC-100-7; and
(f) CC-100-28. (Note: 1 mm = 0.0394 in.)
deflections due to the application of the superimposed load RCA generally resulted in increased beam deflections. Also
were generally underestimated. The analyses were consis- consistent with the experimental results, RCA had greater
tent with the experimental results that increased amounts of effect on the deflections of beams with less cracking. The

184 ACI Structural Journal/January 2018


Table 1—Comparison of analytical results with measured beam deflections

Immediate superimposed load deflection at t = t0, Δsi, mm Immediate plus time-dependent deflection at t = t0 + 119, Δt 0 + 119, mm
Beam ID *
Measured Analytical Analytical/Measured Measured Analytical Analytical/Measured
UT-0-28 0.86 0.74 0.85 5.00 4.27 0.85
UT-0-7 0.74 0.61 0.83 4.62 5.89 1.27
UC-0-28 0.66 0.56 0.85 3.51 2.97 0.85
UC-0-7 0.94 0.64 0.68 5.11 4.04 0.79
CC-0-28 3.15 2.90 0.92 10.19 8.66 0.85
CC-0-7 3.40 3.10 0.91 10.69 10.64 1.00
UT-50-28 0.91 0.74 0.81 5.38 5.08 0.94
UT-50-7 0.94 0.76 0.81 6.96 7.01 1.01
UC-50-28 0.86 0.64 0.74 4.70 3.38 0.72
UC-50-7 0.84 0.64 0.76 5.99 5.36 0.89
CC-50-28 4.93 3.81 0.77 — 10.24 —
CC-50-7 4.14 3.68 0.89 12.90 12.07 0.94
UT-100-28 1.24 0.97 0.78 7.39 8.41 1.14
UT-100-7 1.12 0.94 0.84 — 9.07 —
UC-100-28 0.97 0.84 0.87 5.94 5.92 1.00
UC-100-7 1.27 0.86 0.68 7.62 5.82 0.76
CC-100-28 5.11 6.02 1.18 12.27 13.11 1.07
CC-100-7 4.60 4.67 1.02 14.68 14.38 0.98
*
Beam ID descriptors (Knaack and Kurama 2013b, 2015b): Left descriptors—UT is for beams with bottom (tension) longitudinal reinforcement only and subjected to superim-
posed load designed not to induce immediate cracking; UC is for beams with bottom (tension), top (compression), and transverse reinforcement and subjected to superimposed
load designed not to induce immediate cracking; CC is for beams with bottom, top, and transverse reinforcement and subjected to greater superimposed load to induce immediate
cracking; middle descriptors—0, 50, 100 for R of 0%, 50%, 100%, respectively; right descriptors—28 and 7 for t0 of 28 and 7 days, respectively.
Note: 1 mm = 0.0394 in.

analytical models for Beams UT-0-7 and UC-100-7 were making the prediction of this small amount of cracking
the worst performing among the UT and UC series beams, particularly difficult.
respectively, which may have occurred because of the To evaluate the predictions for local beam behavior,
lack of accompanying creep and shrinkage test data. The comparisons between the measured total strain diagrams
worst performing model from the CC series was for Beam for the CC series beams and the analytical model results
CC-0-28. Companion creep and shrinkage test data were are shown in Fig. 10. The measured total strain trends
available for this beam; however, as a potential source of compared reasonably well with the plane sections assump-
discrepancy, the measured concrete tension strength from the tion used in the analytical models. Note that the analytical
accompanying modulus of rupture tests was disproportion- strain diagrams were zeroed at time t = t0 so that they could
ately large for the measured concrete compression strength. be directly compared with the measurements. The model
The test beam demonstrated a greater level of cracking was generally capable of predicting the downward shift in
(and therefore greater displacements) than the analytical the neutral axis as a result of creep and shrinkage strains
prediction based on the measured (disproportionately large) over time, with more downward shift for beams loaded at a
concrete tension strength. younger age and for beams with increased amounts of RCA.
Although the UC series beams were loaded with a However, even for the initial placement of superimposed
smaller superimposed load to avoid initial cracking, there load at t = t0, the tension strains were significantly underesti-
was still some time-dependent cracking (as induced by mated due to the inability of the model to accurately predict
creep and shrinkage strains) due to the stress redistribu- the amount of cracking. This finding is consistent with the
tions between concrete and reinforcing steel. The analytical beam deflection comparisons in Fig. 9, which show that the
simulations predicted this cracking in all of the UC series initial deflections of the analytical models under superim-
beams, except for Beams UC-0-28 and UC-0-7. Increased posed load were generally smaller than the measured deflec-
creep and shrinkage strains generally caused more cracking tions. The discrepancies in the time-varying deflections
in beams with higher amounts of RCA (that is, higher R). under sustained loading were also likely affected by the
Thus, the extent of creep/shrinkage-induced cracking in limitations of the analytical model to accurately predict the
beams with R = 0% (for example, Beams UC-0-7 and amount of cracking.
UC-0-28) was small as compared to beams with R = 100%, Figure 11 shows the change in the beam neutral axis
depth, c (normalized with respect to the total beam depth,

ACI Structural Journal/January 2018 185


Fig. 10—Total strain diagrams for CC series test beams: (a) CC-0-7; (b) CC-0-28; (c) CC-50-7;
(d) CC-50-28; (e) CC-100-7;
and (f) CC-100-28 (compression negative). (Note: 1 mm = 0.0394 in.)

Fig. 11—Neutral axis depths for CC series test beams: (a) CC-0-7; (b) CC-0-28;
(c) CC-50-7; (d) CC-50-28; (e) CC-100-7;
and (f) CC-100-28.
h) over time, based on the strain diagrams in Fig. 10. The measured strain diagrams, a linear regression analysis was
neutral axis depth was determined as the distance from the conducted on the group of strain measurements taken at a
top of the beam to the location of zero total strain. For the given time to approximately determine the zero intercept of

186 ACI Structural Journal/January 2018


Fig. 12—Midspan stress and strain results for Beam CC-50-7: (a) top fiber stress; (b) top fiber strain components; (c) strain
diagram after loading; and (d) strain diagram prior to unloading (compression negative).
the strains. The results further reinforce the trends discussed analysis. Looking at Fig. 12(b) for the extreme compression
previously that the neutral axis depth increased over time fiber strain components at midspan, it can be seen that the
as a result of the creep and shrinkage deformations of the shrinkage strain was greater than the creep strain before the
concrete. Because the TDConcrete models generally under- application of the superimposed load and was similar to the
estimated the amount of cracking in the beams, the neutral creep strain during the later stages of loading.
axis depth predictions were consistently greater than the The corresponding strain distributions over the beam
measured results. The discrepancies were smaller for depth at midspan are shown in Fig. 12(c) and 12(d) for
beams with R = 100%, which may have been because of times immediately following the application of the super-
the increased cracking in these beams, making the accurate imposed load at t = t0 = 7 days and immediately prior to the
prediction of the extent of this cracking less important. removal of the load at t = 126 days, respectively. The results
Although the companion creep and shrinkage cylinders again show that the shrinkage strain component had a very
described in Knaack and Kurama (2013b, 2015a) showed large effect on the strains over the beam cross section. With
significantly greater creep strains than shrinkage strains, this significant effect of shrinkage on the concrete defor-
the components of the total strain from the beam analyses mations, the assumption of uniform shrinkage throughout
showed that the shrinkage strains had a very significant effect the concrete volume and the uncertainty/variability in the
on the beam deflections, which was a finding that could not measured shrinkage data from the accompanying concrete
be obtained from the experimental data (it was not possible cylinders could have affected the deflection predictions
to separate the measured beam deflections into creep and of the beam specimens. However, considering that a good
shrinkage components). This is because the companion level of agreement was obtained between the analytical and
creep cylinders were loaded to 40% of the concrete compres- measured beam deflections in Fig. 7 to 9, it is concluded
sion strength fc′, while the compression stresses in the beams that the models still provided reasonable estimates of the
generally did not reach this level. For example, Fig. 12(a) shrinkage strains.
shows the extreme (that is, top) concrete compression fiber
stress at the midspan of beam CC-50-7 over the duration Monte Carlo analysis of beam deflections
of the analysis. A maximum concrete compression stress As stated previously, the concrete property inputs in
of approximately 0.22fc′ was reached at the initial place- the TDConcrete analyses of the beams were from average
ment of the superimposed load at t0 = 7 days, with a signif- companion material test data. Because there was significant
icant reduction in concrete stress (due to redistribution of variability in this material test data, the average concrete
stresses into the reinforcing steel) at later times during the properties used to model the beams could have been signifi-

ACI Structural Journal/January 2018 187


Fig. 13—Monte Carlo simulation results: (a) UT-0-7; (b) UT-100-28; (c) UC-50-7; 
(d) UC-100-7; (e) CC-0-28; and (f)
CC-100-28.
cantly different from the actual in-place beam properties. To through these validations, it is concluded that the new time-
investigate this effect, a Monte Carlo study was conducted dependent concrete material model, TDConcrete, and the Creep
by assuming the following concrete properties to be random Analysis developed and implemented in OpenSees can be used
variables: tension strength ft′, secant modulus of elasticity to predict the time-dependent behavior of reinforced concrete
Ec, ultimate creep strain ϵcr,u, and ultimate shrinkage strain structural members, including structures with recycled concrete
ϵsh,u. All other concrete properties and all steel properties aggregates, in a variety of loading scenarios.
were still based on the average measured values from each
set of companion material test data. The mean and standard SUMMARY AND CONCLUSIONS
deviation for the selected random variables were also deter- As a means of analyzing the behavior of reinforced
mined from the material test data for each beam (Knaack concrete structures subjected to sustained service-load condi-
and Kurama 2013b). A Kolmogorov-Smirnov (K-S) test tions, a new time-dependent concrete fiber material model
(Mendenhall and Sincich 2016) was conducted to ensure the was implemented in the open-source OpenSees structural
normality of each data set (ft′, Ec, ϵcr,u, and ϵsh,u). analysis framework. The model was validated using data
By varying ft′, Ec, ϵcr,u, and ϵsh,u within their probability from previous research as well as experimental data from
distributions based on the calculated mean and standard a series of RCA concrete beam specimens. The important
deviation values, each test beam was simulated 200 times findings from the investigation are summarized as follows.
to estimate the range of possible deflections. The results 1. The new time-dependent concrete material model was
of these simulations for six selected beams are shown in able to predict the long-term service-load deformations
Fig.  13, along with the respective measured deflections. of reinforced concrete structural members in a variety of
These beams are among the ones with the greatest discrep- loading scenarios. The initial (instantaneous) deflections
ancies as compared to the original analytical models. The were generally underestimated because of underestimations
results show that except for Beam UT-0-7, the measured in the extent of cracking.
beam deflections fell within the bands created by the Monte 2. The time-dependent concrete material model was also
Carlo simulations, thus supporting the hypothesis that the able to predict the downward shift in the beam neutral axis
inherent variability in the concrete material properties could over time as a result of creep and shrinkage strains. However,
have contributed significantly to the differences between the even for the initial load stage, the tension strains were under-
analytical and measured results in Fig. 7 through 9. As stated estimated due to the inability of the model to accurately
previously, there was no creep or shrinkage data available predict the amount of cracking.
for Beam UT-0-7 and the use of creep and shrinkage data 3. The Monte Carlo investigations showed that the
measured for other beams was the best option available, likely inherent variability in the concrete material properties may
resulting in the greater deflection discrepancies. Overall,

188 ACI Structural Journal/January 2018


have contributed significantly to the differences between the Bažant, Z., and Wu, S., 1973, “Dirichlet Series Creep Function for Aging
Concrete,” Journal of the Engineering Mechanics Division, ASCE, V. 99,
analytical and measured beam deflections. No. 4, pp. 367-387.
4. The analyses were consistent with the experimental CEN, 2004, “Eurocode 2: Design of Concrete Structures – Part 1-1:
results that increased amounts of RCA resulted in increased General Rules and Rules for Buildings (EN 1992-1-1),” Comité Européen
de Normalisation, Brussels, Belgium, 225 pp.
beam deflections. Also consistent with the experimental Choi, W., and Yun, H., 2013, “Long-Term Deflection and Flexural
results, RCA had a greater effect on the deflections of beams Behavior of Reinforced Concrete Beams with Recycled Aggregate,”
with less cracking. Materials & Design, V. 51, pp. 742-750. doi: 10.1016/j.matdes.2013.04.044
Chong, K. T.; Foster, S. J.; and Gilbert, R. I., 2008, “Time-Depen-
5. The time-dependent material model showed that the dent Modelling of RC Structures Using the Cracked Membrane Model
shrinkage strains had a large effect on the long-term deflec- and Solidification Theory,” Computers & Structures, V. 86, No. 11-12,
tions of the beams that were tested, which was a finding that pp. 1305-1317. doi: 10.1016/j.compstruc.2007.08.005
de Borst, R., and van den Boogaard, A., 1994, “Finite-Element Modeling
could not be obtained from the experimental measurements. of Deformation and Cracking in Early-Age Concrete,” Journal of Engi-
6. While further research is certainly needed, a prelimi- neering Mechanics, ASCE, V. 120, No. 12, pp. 2519-2534. doi: 10.1061/
nary expectation from this paper is that RCA concrete may (ASCE)0733-9399(1994)120:12(2519)
Espion, B., 1988, “Long-Term Sustained Loading Tests on Reinforced
be suitable for use in structural applications as long as tools Concrete Beams: A Selected Data Base,” Bulletin du Service Génie, Civil
are available to quantify its effects, such as the increased No. 88-1, Université Libre de Bruxelles.
service-load deflections investigated in this paper. Ghali, A., 1993, “Deflection of Reinforced Concrete Members: A Critical
Review,” ACI Structural Journal, V. 90, No. 4, July-Aug., pp. 364-373.
7. While the paper focuses on applications with RCA Gilbert, I., and Ranzi, G., 2011, Time-Dependent Behaviour of Concrete
concrete, the new modeling capability in OpenSees is Structures, Spon Press, New York.
a general purpose tool that can be used to analyze the Gilbert, R. I., 1999, “Deflection Calculation for Reinforced Concrete
Structures – Why We Sometimes Get it Wrong,” ACI Structural Journal,
time-dependent axial-flexural deformations of conventional V. 96, No. 6, Nov.-Dec., pp. 1027-1032.
reinforced concrete structures as well. Jiang, H., and Kurama, Y., 2010, “Analytical Modeling of Medium-Rise
Reinforced Concrete Shear Walls,” ACI Structural Journal, V. 107, No. 4,
July-Aug., pp. 400-410.
AUTHOR BIOS Kabir, A., 1977, “Nonlinear Analysis of Reinforced Concrete Panels,
ACI member Adam M. Knaack is a Project Engineer at Schaefer in Cincin-
Slabs, and Shells for Time-Dependent Effects,” PhD dissertation, Univer-
nati, OH. He received his MS and PhD in civil engineering from the Univer-
sity of California, Berkeley, Berkeley, CA.
sity of Notre Dame, Notre Dame, IN. He is a member of ACI Committee 435,
Kaklauskas, G., 2004, “Flexural Layered Deformational Model of Rein-
Deflection of Concrete Building Structures, and Joint ACI-ASCE Committee
forced Concrete Members,” Magazine of Concrete Research, V. 56, No. 10,
421, Design of Reinforced Concrete Slabs. His research interests include
pp. 575-584. doi: 10.1680/macr.2004.56.10.575
the design of concrete structures and use of recycled materials in concrete.
Kawano, A., and Warner, R., 1996, “Model Formulations for
Numerical Creep Calculations for Concrete,” Journal of Struc-
ACI member Yahya C. Kurama is a Professor of civil engineering at
tural Engineering, ASCE, V. 122, No. 3, pp. 284-290. doi: 10.1061/
the University of Notre Dame. He is a member of ACI Committees 374,
(ASCE)0733-9445(1996)122:3(284)
Performance-Based Seismic Design of Concrete Buildings, and 555,
Knaack, A., and Kurama, Y., 2013a, “Design of Concrete Mixtures with
Concrete with Recycled Materials, and Joint ACI-ASCE Committee 550,
Recycled Concrete Aggregates,” ACI Materials Journal, V. 110, No. 5,
Precast Concrete Structures. His research interests include the behavior
Sept.-Oct., pp. 483-493.
and design of reinforced concrete structures under extreme loading from
Knaack, A., and Kurama, Y., 2013b, “Sustainable Concrete Structures
earthquake and fire, sustainable construction, and precast/prestressed
using Recycled Concrete Aggregate: Short-Term and Long-Term Behavior
concrete structures.
Considering Material Variability,” Structural Engineering Research Report
#NDSE-2013-01, University of Notre Dame, Notre Dame, IN.
ACKNOWLEDGMENTS Knaack, A., and Kurama, Y., 2015a, “Creep and Shrinkage of Normal
Partial funding for this work was provided by the National Science Foun- Strength Concrete with Recycled Concrete Aggregates,” ACI Materials
dation (NSF) under Grant No. CMMI 1436758. The support of current NSF Journal, V. 112, No. 3, May-June, pp. 451-462. doi: 10.14359/51687392
Program Director Y. G. Hsuan and former Director K. I. Mehta is grate- Knaack, A., and Kurama, Y., 2015b, “Sustained Service-Load Behavior
fully acknowledged. The authors thank D. Atkinson of Concrete Recycling of Concrete Beams with Recycled Concrete Aggregates,” ACI Structural
Center, South Bend, IN, and M. Zeltwanger of American Mobile Aggregate Journal, V. 112, No. 5, Sept.-Oct., pp. 565-577.
Crushing, Nappanee, IN, for their help in acquiring the RCA. Additional Lapko, A., and Grygo, R., 2010, “Long Term Deflection of Recycled
materials were provided by Buzzi Unicem, Sika Corporation, and Transit Aggregate Concrete (RAC) Beams Made of Recycled Concrete,” 10th
Mix South Bend. Any opinions, findings, conclusions, and/or recommenda- International Conference: Modern Building Materials, Structures and
tions in the paper are those of the authors and do not necessarily represent Techniques, Vilnius, Lithuania.
the views of the individuals or organizations acknowledged. Li, X., 2009, “Recycling and Reuse of Waste Concrete in China Part II:
Structural Behavior of Recycled Aggregate Concrete and Engineering
Applications,” Resources, Conservation and Recycling, V. 53, No. 3,
REFERENCES pp. 107-112. doi: 10.1016/j.resconrec.2008.11.005
ACI Committee 209, 1992, “Prediction of Creep, Shrinkage, and
Lou, T. J.; Lopes, S. M. R.; and Lopes, A. V., 2016, “Time-Depen-
Temperature Effects in Concrete Structures (ACI 209R-92),” American
dent Behavior of Concrete Beams Prestressed with Bonded AFRP
Concrete Institute, Farmington Hills, MI, 47 pp.
Tendons,” Composites. Part B, Engineering, V. 97, pp. 1-8. doi: 10.1016/j.
ACI Committee 318, 2014, “Building Code Requirements for Struc-
compositesb.2016.04.070
tural Concrete (ACI 318-14) and Commentary (ACI 318R-14),” American
Maekawa, K.; Soltani, M.; Ishida, T.; and Itoyama, Y., 2006, “Time-
Concrete Institute, Farmington Hills, MI, 519 pp.
Dependent Space-Averaged Constitutive Modeling of Cracked
Ajdukiewicz, A., and Kliszczewicz, A., 2011, “Long-Term Behavior of
Reinforced Concrete Subjected to Shrinkage and Sustained Loads,” Journal
Reinforced-Concrete Beams and Columns Made of Recycled Aggregate
of Advanced Concrete Technology, V. 4, No. 1, pp. 193-207. doi: 10.3151/
Concrete,” fib Symposium, Prague, Czech Republic, pp. 479-482.
jact.4.193
Bacinskas, D.; Kaklauskas, G.; Gribniak, V.; Sung, W. P.; and Shih,
Malešev, M.; Radonjanin, V.; and Marinkovic, S., 2010, “Recycled
M. H., 2012, “Layer Model for Long-term Deflection Analysis of Cracked
Concrete as Aggregate for Structural Concrete Production,” Sustainability,
Reinforced Concrete Bending Members,” Mechanics of Time-Dependent
V. 2, No. 5, pp. 1204-1225. doi: 10.3390/su2051204
Materials, V. 16, No. 2, pp. 117-127. doi: 10.1007/s11043-011-9138-9
Marí, A. R.; Bairan, J. M.; and Duarte, N., 2010, “Long-Term Deflections in
Balevičius, R., 2010, “An Average Stress Strain Approach to Creep
Cracked Reinforced Concrete Flexural Members,” Engineering Structures,
Analysis of RC Uncracked Elements,” Mechanics of Time-Dependent
V. 32, No. 3, pp. 829-842. doi: 10.1016/j.engstruct.2009.12.009
Materials, V. 14, No. 1, pp. 69-89. doi: 10.1007/s11043-009-9093-x
Maruyama, I.; Sogo, M.; Sogabe, T.; Sato, R.; and Kawai, K., 2004,
“Flexural Properties of Reinforced Recycled Concrete Beams,” Interna-

ACI Structural Journal/January 2018 189


tional RILEM Conference on the Use of Recycled Materials in Building Scanlon, A., Murray, D.W., 1974, “Time-Dependent Reinforced-
Structures, Barcelona, Spain, pp. 526-535. Concrete Slab Deflections,” Journal of the Structural Division, ASCE,
Mazzotti, C., and Savoia, M., 2009, “Long-Term Deflection of Rein- V. 100, No. 9, pp. 1911-1924.
forced Self-Consolidating Concrete Beams,” ACI Structural Journal, Sousa, C.; Sousa, H.; Neves, A. S.; and Figueiras, J., 2012, “Numer-
V. 106, No. 6, Nov.-Dec., pp. 772-781. ical Evaluation of the Long-Term Behavior of Precast Continuous Bridge
Mendenhall, W., and Sincich, T., 2016, Statistics for Engineering and Decks,” Journal of Bridge Engineering, ASCE, V. 17, No. 1, pp. 89-96. doi:
the Sciences, sixth edition, CRC Press, Taylor & Francis Group, LLC, Boca 10.1061/(ASCE)BE.1943-5592.0000233
Raton, FL. Tamai, S.; Shima, H.; Izumo, J.; and Okamura, H., 1988, “Average
Mercan, B.; Schultz, A. E.; Stolarski, H. K.; and Magana, R. A., Stress-Strain Relationship in Post Yield Range of Steel Bar in Concrete,”
2013, “Long-Term Lateral Deflection of Precast, Prestressed Concrete Concrete Library of JSCE, V. 11, pp. 117-129.
Spandrel Beams,” PCI Journal, V. 58, pp. 93-115. doi: 10.15554/ Taucer, F.; Spacone, E.; and Filippou, F., 1991, “A Fiber Beam-Column
pcij.09012013.93.115 Element for Seismic Response Analysis of Reinforced Concrete Struc-
Neville, A., 2011, Properties of Concrete, fifth edition, Prentice Hall, tures,” Report No. UCB/EERC-91/17, Earthquake Engineering Research
Upper Saddle River, NJ. Center, University of California, Berkeley, Berkeley, CA.
Nie, J., and Cai, C. S., 2000, “Deflection of Cracked RC Beams under Taylor, R.; Pister, K.; and Goudreau, G., 1970, “Thermomechanical
Sustained Loading,” Journal of Structural Engineering, ASCE, V. 126, Analysis of Viscoelastic Solids,” International Journal for Numer-
No. 6, pp. 708-716. doi: 10.1061/(ASCE)0733-9445(2000)126:6(708) ical Methods in Engineering, V. 2, No. 1, pp. 45-59. doi: 10.1002/
OpenSees, 2006, “The Open System for Earthquake Engineering Simu- nme.1620020106
lation,” Pacific Earthquake Engineering Research Center, Berkeley, CA, Teng, S., and Branson, D., 1993, “Initial and Time-Dependent Defor-
http://opensees.berkeley.edu. (last accessed Nov. 29, 2017) mation of Progressively Cracking Nonprestressed and Partially Prestressed
Rodriguez-Gutierrez, J. A., and Aristizabal-Ochoa, J. D., 2007, “Short- Concrete Beams,” ACI Structural Journal, V. 90, No. 5, Sept.-Oct.,
and Long-Term Deflections in Reinforced, Prestressed, and Composite pp. 480-488.
Concrete Beams,” Journal of Structural Engineering, ASCE, V. 133, No. 4, Washa, G., and Fluck, P., 1952, “Effect of Compressive Reinforcement
pp. 495-506. doi: 10.1061/(ASCE)0733-9445(2007)133:4(495) on the Plastic Flow of Reinforced Concrete Beams,” ACI Journal Proceed-
Sato, R., Maruyama, I., Sogabe, T., Sogo, M., 2007, “Flexural Behavior ings, V. 49, No. 10, Oct., pp. 89-108.
of Reinforced Recycled Concrete Beams,” Journal of Advanced Concrete Zienkiewicz, O. C.; Watson, M.; and King, I. P., 1968, “A Numer-
Technology, V. 5, No.1, pp. 43-61. ical Method of Visco-Elastic Stress Analysis,” International
Journal of Mechanical Sciences, V. 10, No. 10, pp. 807-827. doi:
10.1016/0020-7403(68)90022-2

190 ACI Structural Journal/January 2018


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 115-S15

Carbon Fiber-Reinforced Polymer-Strengthened Reinforced


Concrete Beams Subjected to Differential Settlement
by Yail J. Kim and Aiham Al-Kubaisi
This paper presents the behavior of reinforced concrete beams grooves of a concrete soffit and adhesively bonded (near-sur-
strengthened with carbon fiber-reinforced polymer (CFRP) face-mounted [NSM] strengthening). These are well accepted
composites subjected to differential settlement. Two types of in the rehabilitation community and have numerous advan-
strengthening schemes—externally bonded (EB) and near-surface- tages: durable performance; high strength and modulus;
mounted (NSM) CFRP—are tested with various support condi-
reduced maintenance costs; and prompt implementation with
tions, and their performance is compared against unstrengthened
minimal disruption to users.6,7 Extensive efforts have been
counterparts. The properties of soils are identified by sieve analysis
and load-settlement relationships, including a grain distribution made to examine the performance of CFRP-strengthened
and subgrade reaction moduli with drained and undrained condi- concrete members in various detrimental environments such
tions. The post-yield response and energy dissipation characteris- as fatigue,8 aggressive weather,9 earthquake,10 and fire.11
tics of the beams are influenced by the differential settlement. The However, little is known about the implications of differential
ductility of the strengthened beams is less susceptible to settlement settlement for CFRP-strengthened members. Accordingly,
in comparison with the unstrengthened ones. Regardless of support existing design guidelines do not offer any technical discus-
conditions, the NSM beams exhibit higher ductility than the EB sions or provisions related to this potential hazard. Given that
beams. The development of curvature along the loading span is differential settlement can affect the behavior of the strength-
related to the extent of the differential settlement, which is associ- ened structures, the consequences of such settlement should
ated with the occurrence of CFRP debonding and failure modes.
be elucidated to advance the state of the art of CFRP applica-
Functional requirements for the strengthened beams are evaluated
tions and, eventually, to develop design guidelines.
based on angular distortion limits. Moment redistribution mech-
anisms induced by the differential settlement are elaborated by In this paper, reinforced concrete beams strengthened with
analytical modeling in conjunction with parametric investigations. EB and NSM CFRP are tested under various boundary condi-
tions to examine the effects of differential settlement that can
Keywords: carbon fiber-reinforced polymer (CFRP); differential settle- happen after the beams are strengthened. The focus of the
ment; externally bonded; near-surface-mounted (NSM); strengthening. present investigation is on the consequences of vertical settle-
ment of soils supporting the flexural members, rather than
INTRODUCTION lateral soil movement, including ductility, strain development,
The condition of soil strata is important for structural failure characteristics, angular distortion, and moment redis-
design because their deformation can result in unfavorable tribution. Salient parameters are identified and further used to
distress—namely, deflections and stresses. Excessive settle- analytically expand the experimental findings.
ment degrades the functionality of constructed facilities from
serviceability and strength standpoints. Differential settle- RESEARCH SIGNIFICANCE
ment is critical to structural members compared with uniform The current practice of CFRP strengthening does not
settlement. Differential settlement can also be engaged with consider the movement of foundations, which is crucial for
nonstructural concerns such as unbalanced building floors structures built in regions where unstable soil strata are in
and cracking architectural elements. The economic impact existence. Although ACI 440.2R-0812 is a comprehensive
of settlement-induced structural damage is more than double document on CFRP-based rehabilitation, it does not provide
that associated with natural hazards such as floods, earth- any guidance on strengthened members distressed by differ-
quakes, and hurricanes.1 The movement of soils is typically ential settlement. This research aims to address the identi-
categorized into two stages: immediate settlement and long- fied knowledge gap and to establish a foundation for design
term consolidation. Because the behavior of soils is complex recommendations by understanding the previously unknown
and thus hard to predict (especially when groundwater is soil structure interaction mechanism and its implications.
involved), undesirable settlement often disrupts the opera-
tion of constructed structures. Consequently, the subject of EXPERIMENTAL PROGRAM
differential settlement influencing structural behavior has Materials
been of interest over several decades.2-5 Concrete with a specified compressive strength of fc′
When the flexural capacity of a reinforced concrete = 20 MPa (2900 psi) was mixed with a maximum aggre-
member is insufficient to carry structural loads, it may be
strengthened using carbon fiber-reinforced polymer (CFRP) ACI Structural Journal, V. 115, No. 1, January 2018.
MS No. S-2017-028, doi: 10.14359/51700986, received January 28, 2017, and
composites. Two methods are broadly employed: 1) CFRP reviewed under Institute publication policies. Copyright © 2018, American Concrete
Institute. All rights reserved, including the making of copies unless permission is
sheets are externally bonded along a concrete substrate (EB obtained from the copyright proprietors. Pertinent discussion including author’s
strengthening); and 2) CFRP strips are inserted into precut closure, if any, will be published ten months from this journal’s date if the discussion
is received within four months of the paper’s print publication.

ACI Structural Journal/January 2018 191


Fig. 1—Beam details: (a) cross section and strengthening schemes; (b) steel cage and preparation; and (c) test setup. (Note:
Units in mm; 1 mm = 0.0394 in.)
gate size of 19 mm (0.75 in.). In accordance with ASTM and Af are the elastic modulus and cross-sectional area of
C39/C39M,13 an average compressive strength of 21 MPa the CFRP, respectively) was intended to be the same. The
(3045 psi) was obtained at 28 days. The longitudinal steel substrate of the EB beams was prepared with an electric
reinforcement used was No. 3 deformed bars (As = 71 mm2 grinder to improve bond between the CFRP and the concrete
[0.11 in.2], where As is the cross-sectional area of each bar), (Fig. 1(b)), and a single layer of CFRP sheet was adhered
while the shear reinforcement was No. 2 plain bars (As = using the two-part epoxy. The groove of the NSM beams
32 mm2 [0.05 in.2]) placed at a spacing of 75 mm (3 in.) on (13 mm [0.5 in.] wide by 25 mm [1 in.] deep) was designed
center (Fig. 1(a)). The nominal yield strengths of these bars as per ACI 440.2R-08,12 and was created by mounting
were 414 and 250 MPa (60 and 36 ksi), respectively. Two expanded polystyrene foam underneath the steel cage prior
types of composites were selected as strengthening mate- to casting the concrete. After roughening the groove surfaces
rials: CFRP sheets and strips. The sheets required a wet layup of the hardened beams with the grinder, a CFRP strip was
process with an epoxy adhesive, whereas the strips were inserted with the epoxy adhesive for permanent bonding.
prefabricated. According to the manufacturer, the tensile The EB and NSM beams with a strengthened length of
strength and modulus of the CFRP sheets are 3800 MPa 900 mm (35 in.), as shown in Fig. 1(c), were cured at room
(550 ksi) and 227 GPa (32,900 ksi), respectively, in conjunc- temperature for at least 7 days in compliance with the manu-
tion with an equivalent fiber thickness of 0.165 mm facturer’s recommendation. The identification code of these
(0.0065 in.). The epoxy adhesive comprises a resin and test beams enumerated in Table 1 indicates strengthening
a hardener mixed at a mass ratio of 3:1 to achieve a yield types (UB, EB, and NSM), beam numbers (1 to 12), and
strength of 54 MPa (7800 psi) with an elastic modulus of support conditions (F is fixed, R is rigid, DS is dry sand, and
3 GPa (435 ksi). The tensile strength and modulus of the WS is wet sand). Because the behavior of soil particles is
CFRP strip (16 mm [0.63 in.] wide by 2 mm [0.079 in.] thick) often uncontrollable, three strengthened beams were repeat-
are 2068 MPa (300 ksi) and 124 GPa (18,000 ksi), respec- edly tested per category.
tively. As shown in Fig. 1(b), sand was sieved to determine a
grain distribution and was oven-dried to eliminate moisture Soil test
(details to be discussed). Two kinds of soils were tested: dry sand and wet sand.
The aforementioned oven-dried sand represented the
Beam details dry sand category. For the wet sand, an experimental
Concrete beams were cast with singly and doubly rein- parametric study was conducted to identify a critical
forced sections for the hinged and fixed supports, as shown soil condition when water was involved, which led to
in Fig. 1(a). The top steel bars were sufficiently extended a maximum differential settlement effect. Three water
over 725 mm (29 in.) to preclude the premature cracking contents of 15%, 20%, and 25% of the soil mass were
failure of the beam in negative bending (the inflection point taken into account in drained and undrained conditions
was 538 mm [21 in.] from the fixed end). The beams were (Fig. 2(a)). A load cell and two linear potentiometers
classified into three groups, depending upon strengthening were employed to establish a relationship between the
schemes: unstrengthened (UB); externally-bonded CFRP applied pressure and soil settlement, so that a modulus of
sheets (EB); and near-surface-mounted CFRP strips (NSM). subgrade reaction (representing the stiffness of each soil
For comparison of the strengthened beams’ behavior, the type) was attained.
axial rigidity of the CFRP sheet and strip (EfAf, where Ef
192 ACI Structural Journal/January 2018
Table 1—Identification of test beams
Py, kN Pu, kN
Beam CFRP Soil Each Avgerage Each Avgerage Failure sequence
UB1-F-R None None 41.5 41.5 49.3 49.3 FC → SC → CC
EB2-F-R CFRP sheet None 45.2 53.8 FC → SC → SCT (H)
EB3-F-R CFRP sheet None 49.9 47.6 56.4 56.4 FC → SC → CD (F)
EB4-F-R CFRP sheet None 47.9 59.1 FC → SC → CD (F)
UB5-F-DS None Dry sand 44.6 44.6 48.5 48.5 FC → SC → CC
EB6-F-DS CFRP sheet Dry sand 53.4 59.7 FC → SC → CD (F)
EB7-F-DS CFRP sheet Dry sand 46.4 49.8 63.2 60.8 FC → SC → SCT (H)
EB8-F-DS CFRP sheet Dry sand 49.7 59.7 FC → SC → SCT (F)
UB9-F-WS None Wet sand 44.9 44.9 54.9 54.9 FC → SC → CC
EB10-F-WS CFRP sheet Wet sand 58.9 71.8 FC → SC → CC
EB11-F-WS CFRP sheet Wet sand 56.1 56.1 63.6 67.0 FC → SC → SCT (F)
EB12-F-WS CFRP sheet Wet sand 53.4 65.7 FC → SC → SCT (F)
NSM2-F-R CFRP strip None 51.9 62.5 FC → SC → CD (H)
NSM3-F-R CFRP strip None 53.7 53.9 67.9 67.9 FC → SC → CD (F)
NSM4-F-R CFRP strip None 56.2 73.4 FC → SC → CD (H)
NSM6-F-DS CFRP strip Dry sand 51.2 71.8 FC → SC → CD (F)
NSM7-F-DS CFRP strip Dry sand 51.4 52.7 61.3 66.3 FC → SC → SCT (F)
NSM8-F-DS CFRP strip Dry sand 55.6 65.9 FC → SC → CD (F)
NSM10-F-WS CFRP strip Wet sand 53.5 66.1 FC → SC → SCT (F)
NSM11-F-WS CFRP strip Wet sand 54.1 55.1 63.4 64.8 FC → SC → SCT (F)
NSM12-F-WS CFRP strip Wet sand 57.9 64.9 FC → SC → SCT (F)

Notes: FC is flexural crack; SC is shear crack; CC is concrete crushing; SCT is shear cracked initiated at CFRP termination; CD is CFRP debonding with concrete cover; (H) is CD
near hinge support; (F) is CD near fixed support; 1 kN = 0.225 kip.

Fig. 2—Test setup: (a) soil test; (b) flexural test (fixed rigid); (c) flexural test (fixed settlement); (d) linear potentiometer; (e) PI
gauges; and (f) saturated soil and linear potentiometer.
Test setup and instrumentation configuration was suitable to examine the effects of differen-
The default boundary condition of the beams was one-end- tial settlement. To avoid beam rotation at the fixed end, two
fixed with a hinge on the other side (Fig. 1(c)). This test hollow square steel sections (100 x 100 mm [4 x 4 in.]) were
positioned, which were connected to the rigid steel support

ACI Structural Journal/January 2018 193


Fig. 3—Soil pressure and settlement: (a) drained sand (D); and (b) undrained sand (UD). (Note: 1 kPa = 0.145 psi; 1 mm =
0.0394 in.)
using high-strength threaded rods (19 mm [0.75 in.] in diam- other cases that were fully drained. The K-value of the 25%
eter). A steel plate (100 mm [4 in.] wide by 5 mm [0.2 in.] thick) water sand was, accordingly, 34% and 40% lower than those
was placed on top of the hinge to mitigate concrete crushing of the sand with 15% and 20% water contents, respectively.
caused by stress concentration. Figure 2(b) exhibits the fixed In the case of the undrained sand (Fig. 3(b)), the specimen
rigid (F-R) condition of an EB beam. For the soil-supported with a 15% water content showed hardening after a pressure
hinges (F-DS and F-WS test groups), a circular steel container of 30 kPa (4 psi), which was not observed in its counterparts
with a diameter of 740 mm (29 in.) was used to accommodate involving 20% and 25% water contents. The decreasing trend
a soil depth of 450 mm (18 in.) without boundary effects, as of the reaction modulus was proportional to the amount of
shown in Fig. 1(c) and 2(c). As will be discussed in a later water kept inside the soil voids. The undrained sand with a
section, the undrained soil strata allowed more deflection than water content of 25% (K = 8.4 MPa/m [30 psi/in.]) exhibited
the drained cases; hence, a membrane layer was placed inside more settlement than others; therefore, this soil condition
the container before pouring the oven-dried sand and water was selected for structural testing. The average modulus of
for the F-WS beams. For each test, the soils were positioned subgrade reaction of the dry sand was determined to be K =
anew in the container. 13.7 MPa/m (49 psi/in.).
A load cell, a linear potentiometer, and two displacement-type
strain transducers (so-called PI gauges) were placed at Load displacement
midspan of each beam to measure the flexural behavior Unstrengthened beams—Figure 4(a) shows the load-
(Fig. 2(d) and (e)). Two linear potentiometers at 180 degrees displacement behavior of the unstrengthened beams subjected
apart were also positioned on top of the hinge support to to various boundary conditions. The response of the beam
monitor soil settlement (Fig. 2(f)). In addition, six strain tested at fixed rigid supports (UB1-F-R) was that of a typical
gauges were bonded along the EB CFRP sheet at a typical under-reinforced concrete beam (yield load = 41.5 kN
distance of 150 mm (6 in.), as illustrated in Fig. 1(c). It [9.3 kip] and ultimate load = 49.3 kN [11.1 kip], as listed in
is worth noting that strain gauges were not bonded to the Table 1). The behavior of the beam with fixed dry soil supports
NSM CFRP strip because the placement of the gages might (UB5-F-DS) was similar to the behavior of UB1-F-R up to
influence the bond between the strip and adhesive inside the a load of 15.8 kN (3.6 kip)—32% of the UB1-F-R beam’s
narrow groove. All beams were monotonically loaded in capacity, beyond which a bifurcation was observed owing
four-point bending to failure, and the behavior was recorded to the onset of soil softening. Although there was no notice-
by a computerized data acquisition system. able difference between the capacities of the UB1-F-R and
UB5-F-DS beams, the displacement of the latter at failure
TEST RESULTS was 33% more than that of the former. Additionally, in
Soil properties comparison with the UB1-F-R beam, the UB5-F-DS beam
The pressure-settlement relationship of the sand is plotted accompanied a shorter range of post-yield displacement
in Fig. 3 (average of the two linear potentiometers is given), until failure. This is explained by the fact that, because of
including grain size variation acquired from the sieve analysis. the settlement of the dry soil stratum, the UB5-F-DS beam
The modulus of subgrade reaction K of each test condition substantially deflected before its yield load; consequently,
was taken within a linear settlement range between 5 and the beam’s available deflection was restricted. The ultimate
10 mm (0.2 and 0.4 in.). For the drained sand (Fig. 3(a)), load and corresponding displacement of the beam with wet
all soil responses were similar up to a pressure of approxi- soil (UB9-F-WS) were higher than those of the other beams.
mately 100 kPa (15 psi), regardless of water content. As the The post-yield plateau of UB9-F-WS was not observed
pressure increased, the soil specimen with a water content of (the wet sand dissipated the applied stresses in the beam),
25% began to deviate from those with lower water contents. which confirms the trend of the foregoing reduced post-yield
The reason is that the soil stratum of the 25% water case displacement range. Energy dissipated until the peak loads
preserved a certain amount of residual water in the voids, of the beams was obtained by numerically integrating the
which allowed additional settlement in comparison with the area under the load-displacement curves (Fig. 4(a), inset).

194 ACI Structural Journal/January 2018


Fig. 4—Load displacement of EB beams: (a) unstrengthened beams; (b) fixed rigid; (c) fixed dry sand; and (d) fixed wet sand.
(Note: 1 kN = 0.225 kip; 1 mm = 0.0394 in.)
The energy of UB5-F-DS and UB9-F-WS was 9.5% and placed with care, their unpredictable nature influenced the
63.6% higher than the energy of UB1-F-R, respectively. energy dissipation of the test beams.
EB beams—The behavior of the EB-strengthened beams NSM beams—Figures 5(a) to 5(c) reveal the behavior of the
is shown in Fig. 4(b) to (d), with a comparison against their beams with NSM CFRP strips. The strengthened beam with
unstrengthened counterparts (for clarity, test results of one fixed rigid supports (NSM4-F-R) exhibited a precipitously
representative beam per category are presented). The post- increasing slope up to the yielding of the steel reinforce-
yield stiffness of EB4-F-R was markedly higher than that ment (Fig. 5(a)), followed by an abrupt load drop at 73.4 kN
of UB1-F-R (Fig. 4(b)), and the failure modes of these (16.5 kip). The average energy of the NSM F-R beams was
beams were distinguishable (that is, brittle for EB4-F-R 4.8% higher than that of the unstrengthened UB1-F-R beam,
and ductile for UB1-F-R, which is typical in this kind of which is attributable to the NSM beams’ increased flexural
comparison). A difference of 31.7% in average energy was capacity despite their brittle failure. The flexure of those with
observed between these two categories (Fig. 4(b), inset). The fixed dry sand and fixed wet sand supports (NSM8-F-DS
responses of EB7-F-DS and UB5-F-DS were analogous to and NSM11-F-WS in Figs. 5(b) and 5(c), respectively) was
those of EB4-F-R and UB1-F-R, as shown in Fig. 4(c), except similar to the case of the NSM4-F-R beam; however, because
for the divergence of the displacement at a lower load of of the soil settlement, the former’s stiffness was less than the
20.5 kN (4.6 kip). The regional peak loads of EB7-F-DS latter’s. The load-carrying capacity of the NSM beams tended
after yielding were attributable to the progression of shear to decrease with an increase in settlement, as shown in
cracks initiated at the CFRP termination point (detailed Table 1. For instance, the average failure loads of the NSM
failure modes are elaborated in a later section). Unlike the F-R, F-DS, and F-WS were 67.9, 66.3, and 64.8 kN (15.3,
rigid support case (F-R), the average energy of the EB beams 14.9, and 14.6 kip), respectively. This capacity-varying
with dry sand (F-DS) was 49.8% higher than that of the propensity was opposite to the trend of the EB beams (that
unstrengthened beam (Fig. 4(c), inset). The soil settlement is, the average capacities of 56.4, 60.8, and 67.0 kN [12.7,
of the EB F-DS beams alongside the improved load-carrying 13.7, and 15.1 kip] with the F-R, F-DS, and F-WS support
capacity owing to CFRP strengthening was responsible for conditions, respectively). It is postulated that CFRP-bonded
such energy dissipation. Figure 4(d) illustrates the load-dis- area (EB versus NSM) influences a stress redistribution of the
placement of an EB beam supported by wet sand. The strengthened beams, among other attributes such as failure
strengthened EB10-F-WS beam failed at a 30.8% higher modes, which can alter the capacity of the beams.
load than the unstrengthened one in conjunction with more
displacement. The EB F-WS beams revealed almost double Ductility
energy compared with the UB F-WS beam: 1443 kN·mm The ductility index of the tested beams (DI) was calcu-
(12.8 kip-in.) versus 723 kN·mm (6.4 kip-in.), respectively, lated using Eq. (1) and is summarized in Fig. 5(d)
on average. It is worth noting that, although the soils were

ACI Structural Journal/January 2018 195


Fig. 5—Load displacement of NSM beams: (a) fixed rigid; (b) fixed dry sand; (c) fixed wet sand; and (d) ductility comparison.
(Note: 1 kN = 0.225 kip; 1 mm = 0.0394 in.)

Fig. 6—Comparison of tensile strain at level of reinforcing bar: (a) EB beams; and (b) NSM beams.

Eu as significant as in the unstrengthened beams (Fig. 5(d)).


DI = (1) The ductility indexes of the EB beams with dry sand (F-DS)
Ey
and wet sand (F-WS) were, respectively, 23% and 7% lower
where Eu and Ey are the energy values dissipated up to the than that of the rigid-support beam (F-R). Although similar
ultimate and yield loads of the beam, respectively. The responses were observed, the NSM beams maintained
unstrengthened beam with fixed rigid supports (UB-F-R) consistently higher ductility relative to the EB beams in all
revealed a ductility index of 4.8, which was higher than the circumstances, owing to the NSM CFRP’s enhanced bond
strengthened ones (EB-F-R and NSM-F-R showed DI = 2.2 performance until failure.
and 2.4, respectively). This comparison verifies the well-
known fact that CFRP strengthening reduces the member’s Load strain
ductility.14 When differential settlement was involved, As shown in Fig. 6, where the tensile strain of selected
ductility of the unstrengthened beams dwindled down beams at midspan is summarized, the strain development
to DI = 1.5 and 1.9 for UB-F-DS and UB-F-WS, respec- was controlled by boundary conditions. The general trend
tively. The soil settlement affected the usable deflection of found was that the slope of the load-strain curve became
the beams after yielding of the steel, as discussed earlier, stiffer with an increase in settlement. Namely, the degree of
thereby decreasing the Eu component of Eq. (1). The effect bending in the vicinity of the maximum positive moment
of differential settlement in the strengthened beams was not zone was alleviated. The reason is that the soil-side portion

196 ACI Structural Journal/January 2018


Fig. 7—Soil pressure settlement: (a) EB beams; and (b) NSM beams. (Note: 1 kPa = 0.145 psi; 1 mm = 0.0394 in.)

Fig. 8—CFRP strain of EB beams: (a) fixed rigid (Beam EB2-F-R); (b) fixed dry sand (Beam EB7-F-DS); (c) fixed wet sand
(Beam EB12-F-WS); and (d) comparison at 50% of Beam EB2-F-R’s capacity. (Note: 1 kN = 0.225 kip.)
of the F-DS and F-WS beams shifted down like a rigid where P is the total load applied; a is the distance from the
body, unlike the other portion on the fixed side; therefore, fixed support to the adjacent loading point; b is the distance
the curvature of the beam was not uniform along the span between the two loading points; and L is the span length of
(further discussions are available in subsequent sections). the beam. As far as the EB beams are concerned (Fig. 7(a)),
The load-strain response of the beams appears to be influ- two distinct trends were noticed by soil conditions. The
enced by the strengthening methods; further testing is neces- beams with dry sand exhibited relatively stiffer responses
sary to generate conclusive observations. than those with wet sand, which aligns with the discussion
expounded in the soil property section. For the NSM beams,
Soil settlement a similar trend was recorded with less settlement in the wet
The relationship between the soil pressure and settlement sand condition (Fig. 7(b)). It should be noted that two beams
of the strengthened beams is given in Fig. 7. The soil pres- (EB5-F-DS and NSM10-F-WS) were not included in Fig. 7
sure was calculated using the reaction force at the hinged because the locally disturbed sand caused the unbalanced
support, RH (Eq. (2)), divided by the contact area to the soil displacement of the supporting block, leading to an uplift
under the steel support (Aca = 0.09 m2 [140 in.2]) measurement of the soil settlement.

P  3a (a + b)  CFRP strain
RH = 1 −  (2)
2 2L 2  Figure 8 plots strain development along the CFRP sheets
toward the hinged support (strains to the fixed support side

ACI Structural Journal/January 2018 197


Fig. 9—CFRP strain profile of EB beams: (a) fixed rigid (Beam EB2-F-R); (b) fixed dry sand (Beam EB7-F-DS); and (c) fixed
wet sand (Beam EB12-F-WS). (Note: 1 mm = 0.0394 in.)

Fig. 10—Failure mode of unstrengthened beams: (a) fixed rigid (UB1-F-R); and (b) fixed wet sand (UB5-F-WS).
are available in Fig. 9). For the fixed rigid beam (EB2-F-R in beam (F-R), respectively, because of the reduced curvature
Fig. 8(a)), the strains of Gauges H2 and H3 were similar up effect in tandem with the rigid-body-like shift. The strain
to a load of 12.1 kN (2.7 kip), whereas the strain of Gauge profiles of the CFRP are shown in Fig. 9. The profiles of the
H1 was 67% lower than that of Gauge H2. This indicates beams with fixed rigid and fixed dry sand were generally
that the shear deformation of the adhesive layer (horizontal) symmetric about midspan (Fig. 9(a) and (b), respectively).
caused by the bending of the beam was yet local. When the By contrast, the profile of the beam with fixed wet sand
beam was further loaded, the slope of Gauge H3 was substan- (Fig. 9(c)) was asymmetric, because the effect of the non-
tially reduced, which means that CFRP debonding occurred uniform curvature was augmented by the wet sand stratum.
near midspan of the beam. Gauge H2 followed the same
trend at a load of 20.4 kN (4.6 kip), owing to the progression Failure mode
of CFRP debonding. When diagonal tension cracks began Figure 10 shows the failure modes of the unstrength-
at 30 kN (6.8 kip) near the CFRP’s termination point and ened beams (only selected beams are presented, while
propagated, the response of Gauge H1 became unstable. The detailed failure information of all other beams is available in
beam with fixed dry sand (EB7-F-DS) revealed a tendency Table 1). The failure sequence of the rigid-support beam
analogous to the fixed rigid beam (EB2-F-R), as shown in was typical: flexural cracking, shear cracking, and concrete
Fig. 8(b); however, the strain development of Gauges H2 crushing (Fig. 10(a)). Although the UB beams tested on soils
and H3 (EB7-F-DS) was more rapid. The initial response of demonstrated the same failure sequence (Table 1), the extent
the beam with fixed wet sand (EB12-F-WS in Fig. 8(c)) was of shear damage was more pronounced (Fig. 10(b)), owing
stiffer than those of the other beams, and the strain develop- to the settlement-induced non-uniform curvature effect (that
ment was slow. In compliance with the rigid-body-like shift is, on account of the rigid-body-like shift, the beam portion
explained previously, the shear span of the beam outside the toward the fixed support experienced more damage relative
two-point loading region toward the soil did not experience to the portion on the soil side). The failure of the strength-
significant bending, corroborated by the low strain values of ened beams is provided in Fig. 11. The EB beam with fixed
Gauges H2 and H1 in Fig. 8(c). As summarized in Fig. 8(d), rigid supports (EB3-F-R; Fig. 11(a)) failed by end-peeling
there was an affinity between the CFRP strain near midspan of the CFRP accompanied by concrete-cover delamina-
(Gauge H3) and the support condition at a typical service tion. The failure of the EB beam with dry sand (EB8-F-DS,
state (for example, 50% of the EB2-F-R beam’s capacity). Fig. 11(b)) was initiated by stress concentration at the CFRP
The CFRP strains decreased when the support became softer: termination point, which was then connected to a large
the strains of the beams with dry sand (F-DS) and wet sand diagonal tension crack, including another crack induced by
(F-WS) were 13% and 71% lower than the strain of the rigid excessive negative bending. The NSM4-F-R beam failed

198 ACI Structural Journal/January 2018


Fig. 11—Failure mode of strengthened beams: (a) fixed rigid (EB3-F-R); (b) fixed dry sand (EB8-F-DS); (c) fixed rigid
(NSM4-F-R); and (d) fixed wet sand (NSM10-F-WS).

Fig. 12—Angular distortion of beams: (a) at angle of 0.0033; and (b) at angle of 0.0067.
by CFRP debonding near the rigid support (Fig. 11(c)), on account of a stress redistribution within the beams, which
the same as the EB3-F-R beam (Fig. 11(a)). Nonetheless, led to the shear-crack-dominated failure.
because of the NSM CFRP’s position inside the beam, the
amount of concrete-cover delamination of NSM4-F-R was Angular distortion
more than that of EB3-F-R. Figure 11(d) reveals the failure The angular distortion (AD) of a beam is traditionally
of the NSM beam with fixed wet sand (NSM10-F-WS), defined as the ratio between differential settlement and span
including damage localization within its shear-span near length. This measurement is useful when evaluating the
the fixed support. Unlike the fixed rigid case (NSM4-F-R), performance of a flexural member with the following limits2:
cover delamination failure was not associated despite stress AD = 0.0033 and 0.0067 for the cracking of non-load-bearing
concentration at the CFRP termination. These observations architectural elements and structural damage in a building,
denote that: 1) regardless of strengthening scheme, the beam respectively. Based on the average differential settlement
failure near the fixed support was dominant compared with of each test category, Fig. 12 shows normalized load ratios
the failure near the soil side at occurrence rates of 24%:86%, (a load level of a beam divided by its ultimate capacity Pu)
respectively, as listed in Table 1; and 2) CFRP debonding corresponding to the angular distortion limits. At the archi-
with concrete-cover delamination was mitigated when the tectural limit of AD = 0.0033, the load ratios of the UB, EB,
strengthened beams were subjected to differential settlement and NSM beams tested with dry sand (DS) were P/Pu = 0.37,

ACI Structural Journal/January 2018 199


Fig. 13—Moment ratio with differential settlement in service condition (UB is unstrengthened beam; CFRP is strengthened
beam): (a) maximum negative moment; and (b) maximum positive moment.
0.38, and 0.24, respectively (Fig. 12(a)). When the support
P M 
condition changed to wet sand (WS), these ratios diminished M cs =  − Fs  a (9)
2 L 
by 73%, 71%, and 29%, respectively. Contrary to the UB
and EB categories in F-DS, the NSM F-DS case reached the
where MF is the fixed support negative moment of the beam;
AD limit earlier at P/Pu = 0.24, and the effect of wet sand
MF0 and MFs are the moments at the fixed support without and
altered this tendency. These observations can be explained
with differential settlement, respectively; Mc is the maximum
by the facts that: 1) the average failure load of the NSM
positive moment along the beam span; Mc0 and Mcs are the
F-DS beams was higher than those of the UB and EB beams
moments induced by the applied load P and the differential
(Table 1), resulting in the comparatively low load ratio; and
settlement ∆s, respectively, at which Mc is obtained; and EcI
2) the NSM beams’ load-carrying capacity decreased by
is the flexural rigidity of the transformed section consisting
3% with a change in soil type from dry sand to wet sand
of the concrete elastic modulus (Ec = 57,000√fc′ in psi [Ec =
(Table 1), which was different from the variation trend of
4700√fc′ in MPa]15) and the moment of inertia at an arbitrary
the other beams, as delineated previously. The load ratios of
loading stage within the elastic domain (gross and cracked).
the UB, EB, and NSM beams with dry sand at the structural
Figure 13 shows the effects of differential settlement on the
damage limit of AD = 0.0067 were P/Pu = 0.54, 0.61, and
service moment of the beams before and after cracking loads.
0.44, respectively, and their reduction rates concerned with
Because the total negative and positive moments (MF and Mc,
wet sand were reasonably preserved, as in the case of AD =
respectively) were normalized by the base components MF0
0.0033 (that is, 61%, 72%, and 20%, respectively). Further
and Mc0, as shown in the equations in Fig. 13, their ratios were
discussions on the relationship between the strengthening
maintained within the two load categories (that is, P ≤ Pcr and
scheme and differential settlement follow in a later section.
Pcr < P ≤ Py, where Pcr and Py are the cracking and yield loads
of the beam, respectively). It is worth again noting that the
Service moment due to differential settlement
transformed sectional properties of the EB and NSM beams
Fundamental structural analysis provides the following
were the same due to the identical axial rigidity of the CFRP
expressions in conjunction with the test configuration illus-
sheet and strip; hence, a unified expression “CFRP” was used
trated in Fig. 1
in Fig. 13. For the negative moment (Fig. 13(a)), the ratio of
the rigid support case (FR) was unity owing to the null contri-
MF = MF0 + MFs (3)
bution of the MFs and Mcs to the total moment. As the support
condition changed from dry sand (F-DS) to wet sand (F-WS),
Mc = Mc0 + Mcs (4)
the moment ratios before cracking increased to 2.4 and 3.3,
respectively. After cracking of the beams, the moment ratio
3Pa (a + b) was influenced by the presence of CFRP, because the contri-
MF0 = (5)
4L bution of the CFRP to the cracked beam section was not negli-
gible, unlike the case of the uncracked section for the gross
3Ec I moment of inertia. The post-cracked moment ratios of the
M Fs = ∆ s (6)
L2 strengthened beams were 8% and 9% higher than those of the
unstrengthened beams for the dry sand and wet sand condi-
RH tions, respectively. For the positive moment (Fig. 13(b)), the
∆s = (7) amounts of the increased moment ratios caused by the settle-
K s Aca
ment were not as high as those of the ratios in the negative
moment. Moreover, after cracking, the strengthened beams
Pa  3a (a + b)  with settlement revealed lower moment ratios than did the
M c0 = 1 −  (8)
2  2 L2  unstrengthened ones, opposite to the negative moment trend.
These facts signify that the positive moment caused by differ-

200 ACI Structural Journal/January 2018


Fig. 14—Effects of properties of cracked beams (UB is unstrengthened beam; CFRP is strengthened beam): (a) concrete
strength; (b) modulus of subgrade reaction; (c) elastic modulus of CFRP; and (d) cross-sectional area of CFRP. (Note:
1 MPa/m = 3.68 psi/in.; 1 MPa = 145 psi; 1 mm2 = 0.0016 in.2.)
ential settlement was redistributed to the negative moment, from the beams’ default subgrade reaction modulus of
when the strengthened beams cracked. It should be noted that K = 8.4 MPa/m (31 psi/in.), the negative moment ratio of
the moment redistribution resulting from differential settle- the unstrengthened (UB) and strengthened (CFRP) beams
ment is different from the conventional redistribution associ- increased from 2.6 to 3.6 and from 2.9 to 3.9, respectively.
ated with the yielding of steel reinforcement. Although the CFRP beam revealed higher moment ratios
than did the UB beam, their response rates were identical.
PARAMETRIC STUDY These increased negative moment ratios imply that, when
Formulation differential settlement takes place, CFRP strengthening of
Several parameters affect the negative and positive stiff beams (or high-strength concrete beams) exacerbates
moments of the one-end-fixed beam subjected to differential the fixed end moment. In contrast, the positive moment
settlement: the compressive strength of concrete, the elastic of the beams was lessened because of the aforementioned
modulus of CFRP, the number of CFRP layers, and the moment redistribution mechanism. Figure 14(b) illustrates
modulus of subgrade reaction. Based on Eq. (3) to (9), the the moment variation of the beams with variable subgrade
effects of these parameters were studied within a practical reaction moduli at a concrete strength of fc′ = 20 MPa
range, as shown in the abscissa of Fig. 14. The configura- (2900 psi). As the reaction moduli rose, the negative
tion of the test beam (Fig. 1) was the default for the present moment ratios were asymptotically reduced toward unity.
parametric investigation into moment responses with and Given that the increased reaction modulus aligns with a
without CFRP strengthening. The applied load P in Fig. 1(c) reduction in differential settlement, the extent of additional
was set to represent a typical service condition (that is, 60% negative moment caused by the settlement decreased and
of the yield load of the control beam, UB1-F-R). Because should disappear if the reaction modulus is significantly high
there was no remarkable difference between the EB- and like a rigid support. On the contrary, the positive moment
NSM-strengthened beams in service as discussed previ- ratios increased with the increased reaction moduli, because
ously, these strengthening methods were not differentiated. the moment redistribution mechanism was diminished. The
contribution of the CFRP’s moduli to the moment ratios is
Effects of parameters provided in Fig. 14(c). When the CFRP became stiffer, both
The effect of compressive concrete strength fc′ is shown the negative and positive moment ratios of the strengthened
in Fig. 14(a). This parameter is engaged with the elastic beams deviated from the invariant ratios of the unstrength-
modulus of the concrete Ec that influences the flexural ened cases: the variation of the negative moment was more
rigidity of the beam (Eq. (5) and (9)). With an increase in pronounced than that of the positive moment. Similar obser-
concrete strength from 20 MPa to 50 MPa (2900 to 7250 psi) vations were noted for the beams strengthened with variable
at a differential settlement of 10.5 mm (0.41 in.) resulting

ACI Structural Journal/January 2018 201


CFRP layers, as shown in Fig. 14(d), except that the CFRP a member of ACI Committees 342, Evaluation of Concrete Bridges and
Bridge Elements; 343, Concrete Bridge Design; and 440, Fiber-Reinforced
effect was augmented relative to the case of Fig. 14(c). Polymer Reinforcement. His research interests include advanced composite
materials for structural rehabilitation; complex systems; uncertainty quan-
SUMMARY AND CONCLUSIONS tification; and science-based structural engineering, including statistical,
interfacial, and quantum physics.
This experimental research has investigated the behavior
of CFRP-strengthened beams under differential settlement, Aiham Al-Kubaisi is a former masters student in the Department of Civil
which was not well understood in the community. The effects Engineering at the University of Colorado Denver. He received his BS in
civil engineering from Nahrain University, Baghdad, Iraq, in 2008. His
of strengthening schemes (EB and NSM) were comparatively research interests include structural strengthening using composite mate-
examined with various support conditions (rigid, dry sand, and rials and soil-structure interaction.
wet sand). Soil testing was conducted to study the influence of
water inside the soil strata with drained and undrained condi- ACKNOWLEDGMENTS
tions. Although the number of test beams was limited, consistent The authors would like to acknowledge financial support from the Higher
Committee for Education Development in Iraq (HCED). Technical contents
observations were made in most cases. An analytical parametric presented in this paper are based on the opinion of the authors, and do not
study was carried out to assess the contribution of typical engi- necessarily represent that of others. Proprietary information such as product
neering parameters affecting the performance of the strengthened names and manufacturers is not provided to avoid commercialism.
beams. A follow-up study may be of interest with various sizes of
CFRP-strengthened structural members and connection details REFERENCES
1. Holtz, W. G., and Hart, S. S., “Home Construction on Shrinking and
subjected to differential settlement. The following is concluded: Swelling Soils,” Special Publication 11, Colorado Geological Survey,
• The amount of water inside the soil strata controlled the 1978, 18 pp.
subgrade reaction modulus, which influenced the extent 2. Skempton, A. W., and MacDonald, D. H., “The Allowable Settlements
of Buildings,” Proceedings — Institution of Civil Engineers, V. 5, No. 6,
of settlement. The available post-yield displacement of 1956, pp. 727-768. doi: 10.1680/ipeds.1956.12202
the unstrengthened beams until failure was reduced by 3. Moulton, L. K.; GangaRao, H. V.; and Halvorsen, G. T. “Tolerable
the occurrence of differential settlement when compared Movement Criteria for Highway Bridges,” Report No. FHWA/RD-85/107,
Federal Highway Administration, Washington, DC, 1985, 109 pp.
with displacement of the beam tested at rigid supports. 4. Hearn, G., and Nordheim, K., “Differential Settlements and Inelastic
Nonetheless, the settled beams showed higher energy Response in Steel Bridge Beams,” Transportation Research Record:
dissipation than the rigid beam. The support conditions Journal of the Transportation Research Board, V. 1633, 1998, pp. 68-73.
doi: 10.3141/1633-09
also affected energy dissipation characteristics of the 5. Hebeler, G. L.; Ospina, R. I.; Howard, D. K.; and Leistikow, R.,
strengthened beams. “Assessment and Remediation of Geotechnical and Structural Deficiencies
• The ductility of the tested beams decreased after CFRP at a 5-Story Hotel Experiencing Differential Settlements,” Proceedings of
the 5th Congress on Forensic Engineering, 2009, pp. 278-287.
strengthening. The variation of ductility in the strength- 6. Bakis, C. E.; Bank, L. C.; Brown, V. L.; Cosenza, E.; Davalos, J. F.;
ened beams was not as susceptible to differential settle- Lesko, J. J.; Machida, A.; Rizkalla, S. H.; and Triantafillou, T. C., “Fiber-
ment as the case of the unstrengthened beams. The Reinforced Polymer Composites for Construction—State-of-the-Art
Review,” Journal of Composites for Construction, ASCE, V. 6, No. 2, 2002,
NSM beams demonstrated higher ductility than the EB pp. 73-87. doi: 10.1061/(ASCE)1090-0268(2002)6:2(73)
beams, irrespective of support condition. 7. Bank, L. C., Composites for Construction, John Wiley & Sons, Inc.,
• Differential settlement caused the rigid-body-like shift of the Hoboken, NJ, 2006, 551 pp.
8. Zhang, W., and Kanakubo, T., “Local Bond Stress-Slip Relationship
beams, which disrupted the formation of uniform curvature between Carbon Fiber-Reinforced Polymer Plates and Concrete under
along the loading span. Accordingly, strain development in Fatigue Loading,” ACI Structural Journal, V. 111, No. 4, July-Aug. 2014,
the positive moment region of the beams was alleviated, pp. 955-966. doi: 10.14359/51686781
9. Deng, J.; Tanner, J. E.; Mukai, D.; Hamilton, H. R.; and Dolan, C. W.,
including the strains along the CFRP sheets. The onset of “Durability Performance of Carbon Fiber-Reinforced Polymer in Repair/
CFRP debonding was mitigated by the settlement. Strengthening of Concrete Beams,” ACI Materials Journal, V. 112, No. 2,
• Owing to the damage localization associated with diag- Mar.-Apr. 2015, pp. 247-257. doi: 10.14359/51687104
10. Parks, J. E.; Brown, D. N.; Ameli, M. J.; and Pantelides, C. P.,
onal tension cracks, most strengthened beams failed on the “Seismic Repair of Severely Damaged Precast Reinforced Concrete Bridge
fixed-end side, rather than on the soil-supported side. As far Columns Connected with Grouted Splice Sleeves,” ACI Structural Journal,
as functional requirements were concerned, the strengthened V. 113, No. 3, May-June 2016, pp. 615-626. doi: 10.14359/51688756
11. Williams, B.; Kodur, V.; Green, M. F.; and Bisby, L., “Fire Endurance
and unstrengthened beams reached the angular distortion of Fiber-Reinforced Polymer Strengthened Concrete T-beams,” ACI Struc-
limits at lower loads because of the differential settlement. tural Journal, V. 105, No. 1, Jan.-Feb. 2008, pp. 60-67.
• CFRP strengthening increased the negative moment of the 12. ACI Committee 440, “Guide for the Design and Construction of
Externally Bonded FRP Systems for Strengthening Concrete Structures
beams when subjected to differential settlement, because (ACI 440.2R-08),” American Concrete Institute, Farmington Hills, MI,
the positive moment of the strengthened beams was 2008, 76 pp.
redistributed to the negative moment. With an increase in 13. ASTM C39/C39M-16, “Standard Test Method for Compressive
Strength of Cylindrical Concrete Specimens,” ASTM International, West
concrete strength, the fixed end moment of the strength- Conshohocken, PA, 2016, 7 pp.
ened beams worsened due to the redistribution mechanism. 14. Spadea, G.; Swamy, R. N.; and Bencardino, F., “Strength and
Ductility of RC Beams Repaired with Bonded CFRP Laminates,” Journal
of Bridge Engineering, ASCE, V. 6, No. 5, 2001, pp. 349-355. doi: 10.1061/
AUTHOR BIOS (ASCE)1084-0702(2001)6:5(349)
Yail J. Kim, FACI, is a Professor in the Department of Civil Engi-
15. ACI Committee 318, “Building Code Requirements for Structural
neering at the University of Colorado Denver, Denver, CO. He is Chair
Concrete (ACI 318-14) and Commentary (ACI 318R-14),” American
of ACI Committee 345, Concrete Bridge Construction, Maintenance, and
Concrete Institute, Farmington Hills, MI, 2014, 520 pp.
Repair; and ACI Subcommittee 440-I, FRP-Prestressed Concrete. He is

202 ACI Structural Journal/January 2018


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 115-S16

Tests of Short Headed Bars with Anchor Reinforcement


Used in Beam-to-Column Joints
by Ján Bujn ák and Matúš Farbák

The proper design and execution of joints is one of the basic


preconditions for efficient functioning of precast concrete frames.
Headed reinforcement bars are an efficient technique to anchor
tensile forces in beam-to-column joints. In comparison to tradi-
tional anchorage techniques that consist in developing tensile bars
in the form of a bend or a hook, headed bars allow for optimiza-
tion of the amount of steel and prevent congestion of reinforcement
in the column. The present paper deals with tests of short headed
bars anchoring tensile forces in compact columns. Practical bene-
fits yielding from the research are shown on examples of industrial
building products.

Keywords: anchor reinforcement; beam-column joint; headed bars; tensile


tests.

INTRODUCTION
A number of design guidelines are currently available to
assess the performance of headed bars used in beam-column
joints. ACI 318-141 allows the assumption that headed bars Fig. 1—Headed bar with anchor reinforcement.
are equivalent to hooks or bends provided that ldt > d/1.5,
and the bars are developed to the side of the column opposite l1 ⋅ π ⋅ d s ⋅ f bd
of the joint. In all other cases (for example, short headed bars N Rd , a = ∑ (1)
n α
with ldt < d/1.5), the resistance of the anchorage against cone
breakout needs to be verified. fib Bulletin 722 introduces an Comprehensive information about the background of
approach similar to ACI 318-141 with a ratio of heff/d > 0.6 this design model may be found in reference published by
required to treat the headed bar in a manner equivalent to a Eligehausen et al.5 The logic and practical implications of
bend or a hook. Eurocode EN 1992-1-13 does not formu- Eq. (1) for the design of headed anchors has been recently
late methods for the assessment of headed bars. Within researched by several authors. Henriques et al.6 tested indi-
current European legislation, the performance of products vidual headed bars and groups of headed bars with embed-
and systems that use headed bars as anchorage elements ment depth heff = 150 mm (5.91 in.) coupled with anchor
is usually assessed by product approvals that demonstrate reinforcement, observing failure loads significantly higher
the conformity of the headed bars to technical specification in comparison to those predicted using Eq. (1). Similar
CEN/TS 1992-4-2,4 where the resistance of the headed bars tests and conclusions have been published by Bujnak et al.7
is typically limited by the concrete failure (for example, for groups of headed bars with anchor reinforcement and
concrete cone failure) even with relatively long headed bars. anchorage depth as short as heff = 65 mm (2.56 in.). From-
Where the tensile capacity of the headed bar is governed knecht et al.8 tested fastening plates with four headed bars,
by the concrete cone breakout, the codes referred to herein heff = 150 mm (5.91 in.), and anchor reinforcement. The
allow use of anchor reinforcement to increase the capacity plates were embedded in columns and loaded by eccentric
of the anchorage. Most typically, anchor reinforcement is shear load. The tests allowed identification of various failure
provided in the form of stirrups that are designed and detailed modes with failure loads four to eight times higher compared
to tie the potential breakaway cone to the rest of the concrete to the characteristic value of resistance determined in accor-
body (Fig. 1). ACI 318-141 allows use of anchor reinforce- dance with CEN/TS 1992-4-2.4 Berger9 developed a new
ment provided that the anchorage length of the stirrup in the design model for anchors with anchor reinforcement. The
concrete cone (equivalent to l1 from Fig. 1) is greater than model was verified using a comprehensive database of tests
8ds or 152 mm (6 in.). The technical specification CEN/TS
ACI Structural Journal, V. 115, No. 1, January 2018.
1992-4-24 limits the capacity of the anchor reinforcement MS No. S-2017-049.R2, doi: 10.14359/51701121, received March 15, 2017, and
with the following design value of resistance reviewed under Institute publication policies. Copyright © 2018, American Concrete
Institute. All rights reserved, including the making of copies unless permission is
obtained from the copyright proprietors. Pertinent discussion including author’s
closure, if any, will be published ten months from this journal’s date if the discussion
is received within four months of the paper’s print publication.

ACI Structural Journal/January 2018 203


Fig. 2—Alternative methods to anchor tensile forces to a
column: (a) bend or hook; (b) short headed bar with anchor
reinforcement; and (c) long headed bar developed to oppo-
site side of column.
Fig. 3—Geometric properties of headed anchor.
available from literature as well as Berger’s own tests. The
RESEARCH SIGNIFICANCE
model formulates the anchorage failure of the anchor rein-
The experimental program has been performed to over-
forcement with a resistance value significantly higher than
come the lack of knowledge about the structural performance
the value provided by Eq. (1). Meanwhile, it also introduces
of deformed short headed bars (heff = 140 mm [3.94  in.])
a maximum resistance that is defined as a factored breakout
with a large diameter (25 mm [0.98 in.]) and short overlap
resistance of concrete
length with anchor reinforcement (l1 = 81 mm [3.18 in.]).
Test series A allowed optimization of the test procedure,
NRd,max = ψ ∙ NRd,c (2)
identifying the minimum amount of reinforcement needed
for a reliable functioning of the test setup. Test series B and
with the coefficient ψ having a maximum value of 2.6. The
C demonstrated the failure loads and modes of short headed
tests of headed bars with anchor reinforcement summarized
bars with anchor reinforcement. A significant difference
herein have been performed using headed anchors made
between the measured value and failure loads predicted by
of smooth, round bars with a maximum diameter 16 mm
available design methods was observed.
(0.63 in.). Except for a handful of tests with shorter embed-
ment depth, the embedment depth of anchor bars used in
EXPERIMENTAL INVESTIGATION
the tests is 150 mm (5.91 in.) or greater. All of the works
The experimental program consisted of three test series,
dealing with the assessment of the current design methods
where the tensile capacity of headed anchors cast in concrete
against results of experimental research conclude that design
columns was tested. Identical anchors with dimensions summa-
methods that use Eq. (1) are extremely conservative.
rized in Fig. 3 were used in all tested specimens. The headed
The use of industrially manufactured building products
bars are manufactured from European deformed bars B500B
might have a significant impact on the efficiency of precast
with characteristic yield strength 500 N/mm2 (72.52 ksi).
concrete construction. Within beam-column joints, such
The headed bars were welded to a distribution plate with
building products (corbels, reinforcing bar couplers, or
a round hole. The distribution plate was connected to a steel
others) carry efforts from the beam and anchor them in the
rod through a hinge. The load was applied to the steel rod
column. To secure the continuity of forces between the beam
by a hollow hydraulic jack (Fig. 4). The load was applied
and the column, the anchorage part of the building products
continuously up to failure and measured by a load cell
is typically manufactured from deformed bars with forms
placed on the top of the hydraulic jack. The displacements
illustrated in Fig. 2: (a) bend developed to the opposite side
were measured using gauges placed between the distribution
of the column; (b) short headed bar with anchor reinforce-
plates and the bottom of the concrete column.
ment; and (c) long headed bar developed to the opposite side
The concrete specimens were reinforced horizontal
of the column.
and transverse reinforcement. Both tests with and without
While options (a) and (c) are in line with current design
anchor reinforcement placed around the headed bar have
codes (for example, ACI 318-141) and their design is thus
been conducted.
quite straightforward, option (b) is preferable from indus-
trial point of view because it allows standardization of the
Test series A
manufacturing process of building products and thus reduces
The target of test series A was to evaluate the amount
their cost for the construction industry. At the same time, the
of horizontal and transverse reinforcement needed for the
available data from literature indicate that the use of CEN/TS
proper execution of test. Indeed, while the bending and shear
1992-4-24 for the design of such short headed anchors leads
reinforcement of the concrete member might positively
to a design that is inefficient from a cost point of view. This is
affect the local performance of the anchorage, it is neces-
what motivates the designers of short headed anchors to look
sary to avoid the global failure of the tested specimens by
for alternative design methods—that is, design by testing.
bending or shear. The geometry and material parameters of

204 ACI Structural Journal/January 2018


Fig. 4—Test setup.

Fig. 5—Geometric parameters of specimens, Series A.


the specimens tested in Series A are summarized in Fig. 5
and Table 1.
The load displacement behavior of Series A is illustrated
in Fig. 6 and the maximum values of load recorded during
the test are in Table 1. The only specimen where the failure
of the anchorage was observed was Specimen 3A (Fig. 7).
The failure of Specimens 2A and 4A was associated with the
development of diagonal cracks starting from the bottom of
the tested specimen (shear failure of the concrete member),
and the failure of Specimen 5A was associated with the
opening of vertical cracks in the top part of the specimen
(bending failure).
Fig. 6—Load-displacement curves of specimens, Series A.

ACI Structural Journal/January 2018 205


Table 1—Test specimens, series A
Concrete beam Beam reinforcement Anchor reinforcement
fc,cube, m, N/mm
2
Bars, mm Stirrups, mm Bars, mm Stirrups, Ntest, kN
Specimen l, mm (in.) b, mm (in.) h, mm (in.) (ksi) (in.) (in.) (in.) mm (in.) (kip)
1A Ø10 (0.39) — 134.5 (30.2)
2A Ø20 (0.79) — 246.4 (55.4)
1500 280 280 41.5 4 Ø10 2 Ø10
3A Ø20 (0.79) Ø8 (0.31) 302.6 (68.0)
(59.06) (11.02) (11.02) (6.02) (0.39) (0.39)
4A Ø12 (0.47) — 187.2 (42.1)
5A Ø12 (0.47) Ø8 (0.31) 213.6 (48.0)
Reinforcement, fy, N/mm2 fu, N/mm2
mm (in.) (ksi) (ksi)
Ø8 (0.31) 609 (88.3) 641 (93.0)
Ø10 (0.39) 532 (77.2) 610 (88.5)
Ø12 (0.47) 544 (78.9) 627 (90.9)
Ø25 (0.98) 555 (80.5) 651 (94.4)

Fig. 7—Specimens 2A, 3A, and 4A at failure load and Specimens 1A and 5A after failure.
Test series B and C reinforcement; all other specimens were equipped with
Based on observations yielding from test series A, the spec- anchor reinforcement.
imens tested in series B and C were equipped with normal The load displacement behavior of Specimens B and C is
reinforcement consisting of four reinforcing bars Ø25 mm illustrated in Fig. 9. The failure of Specimen 1C happened
(0.98 in.) and a shear reinforcement intended to prevent the in the welds between the headed bar and the distribution
shear failure of the specimen outside the anchorage area. plate. In all other specimens with anchor reinforcement, the
Test series B had a concrete with average cubic strength failure was associated with the development of cracks on
(measured on cubes with 150 mm [5.91 in.] edges) fc,cube,m the surface of the concrete specimen around the headed bar
= 40.07 N/mm2 (5.81 ksi); the average cubic strength in test (Fig. 10). Headed bars without anchor reinforcement were
series C was fc,cube,m = 52.5 N/mm2 (7.61 ksi). The dimen- pulled out of concrete (Fig. 11).
sions and parameters of specimens tested in series B and C
are summarized in Fig. 8 and Table 2. The varied parame- EXPERIMENTAL RESULTS AND DISCUSSION
ters were the amount and diameter of anchor reinforcement The direct comparison between tests with anchor rein-
and transverse reinforcing bars placed in the corners of the forcement and reference tests without anchor reinforcement
anchor reinforcement. Specimens 4B and 3C had no anchor indicates the positive effect of the anchor reinforcement
on the performance of the anchorage. Depending on the

206 ACI Structural Journal/January 2018


Fig. 8—Geometric parameters of specimens, Series B and C.

Table 2—Test specimens, Series B and C


Concrete beam Beam reinforcement Anchor reinforcement
fc,cube, m, Bars, mm Stirrups, mm Bars, mm Stirrups, mm Ntest, kN
Specimen l, mm (in.) b, mm (in.) h, mm (in.) N/mm2 (ksi) (in.) (in.) (in.) (in.) (kip)
1B 4 Ø10 (0.39) 2 Ø10 (0.39) 294.9 (66.3)
2B 280 280 4 Ø12 (0.47) 2 Ø10 (0.39) 303.7 (68.3)
3B (11.02) (11.02) 4 Ø10 (0.39) 4 Ø8 (0.31) 330.4 (74.3)
1500 40.07 Ø 25 Ø8
4B (59.06) (5.81) (0.98) (0.31) — — 138.4 (31.1)
5B 380 (14.96) 280 (11.02) 292.8 (65.8)
4 Ø10 (0.39) 2 Ø10 (0.39)
6B 380 (14.96) 380 (14.96) 293.7 (66.0)
Reinforcement, fy, N/mm 2
fu, N/mm 2

mm (in.) (ksi) (ksi)


Ø8 (0.31) 535 (77.6) 693 (100.5)
Ø10 (0.39) 524 (76.0) 617 (89.5)
Ø25 (0.98) 567 (82.2) 657 (95.3)
Beam
Concrete beam reinforcement Anchor reinforcement
fc,cube, m, Bars, mm Stirrups, mm Bars, mm Stirrups, mm Ntest, kN
Specimen l, mm (in.) b, mm (in.) h, mm (in.) N/mm2 (ksi) (in.) (in.) (in.) (in.) (kip)
1C 4 Ø10 (0.39) 2 Ø10 (0.39) 343.4 (77.2)
1500 280 280 52.5 Ø 25 Ø8
2C — 2 Ø10 (0.39) 305.6 (68.7)
(59.06) (11.02) (11.02) (7.61) (0.98) (0.31)
3C — — 167.7 (37.7)
Reinforcement, fy, N/mm2 fu, N/mm2
mm (in.) (ksi) (ksi)
Ø8 (0.31) 547 (79.3) 615 (89.2)
Ø10 (0.39) 549 (79.6) 640 (92.8)
Ø25 (0.98) 541 (78.5) 630 (91.4)

ACI Structural Journal/January 2018 207


amount of steel used (both anchor reinforcement and trans- where fbm = 2.25 ∙ η1 ∙ η2 ∙ fctm.
verse bars), the ratio is between 2.05 and 2.39 (Table 3). The The mean value of ktest = 3.66 with COV = 17.79%. The
direct comparison between tests 1C and 2C also illustrates value ktest is determined as
the beneficial effect of the transverse bars on the anchorage
capacity. This influence is currently neglected by available N test
design models. ktest = (4)
N R , calc
The comparison of the measured and calculated values indi-
cates that the available design models might not allow assess- where
ment of the performance of headed bars to be applied in a real-
istic manner such as the tested configuration. The ACI model Ac , N
is not applicable because the anchorage length of the stirrup in N R , calc = ⋅ (11.9 ⋅ f cm , cube ⋅ hef1.5 ) ⋅ ψ s , N ⋅ ψ re , N ⋅ ψ ec , N (5)
the concrete cone is smaller than the minimum value specified Ac0, N
by the code. The mean value of Ntest/NRm,a = 3.37 with COV =
11.40%, where the calculated value is determined as Ac,N = (3 ∙ heff) ∙ b (6)

l1 ⋅ π ⋅ d s ⋅ f bm A0c,N = (3 ∙ heff)2 (7)


N Rm , a = ∑ (3)
n α
where ψs,N, ψre,N, and ψec,N are factors according to CEN/TS
1992-4-2.4

PRACTICAL IMPLICATIONS
The tests presented herein have been used to optimize the
anchorage part of an innovative corbel system that is used as
a vertical support for reinforced concrete or steel-concrete
composite beams.10 It consists of an anchorage part that is
integrated in a precast concrete column and a steel bracket
that is bolted to the anchorage part after the removal of the
column’s formwork (Fig. 12).
The anchorage part consists of a system of vertical and
horizontal anchorage bars. Variants of the anchorage
part of the corbel used up to now in Europe are shown in
Fig. 13. While the anchorage in Fig. 13(a) complies with
EN 1992-1-1,3 the shape and length of the anchor bars
Fig. 9—Load-displacement curves, series C and B.

Fig. 10—Specimens 1B through 6B at failure load.

208 ACI Structural Journal/January 2018


Table 3—Summary of predictions and experimental results
Anchor r. Concrete
Specimen Ntest, kN (kip) Stirrups, mm (in.) l1, mm (in.) fcm,cube, N/mm (ksi)
2
fctm, N/mm2 (ksi) fbm, N/mm2 (ksi) NRm,a, kN (kip) Ntest/NRm,a
1B 294.9 (66.3) 2 Ø10 (0.39) 81 (3.19) 40.07 (5.81) 2.50 (0.36) 5.62 (0.82) 81.8 (18.4) 3.61
2B 303.7 (68.3) 2 Ø10 (0.39) 81 (3.19) 40.07 (5.81) 2.50 (0.36) 5.62 (0.82) 81.8 (18.4) 3.71
3B 330.4 (74.3) 4 Ø8 (0.31) 78 (3.07) 40.07 (5.81) 2.50 (0.36) 5.62 (0.82) 126.0 (28.3) 2.62
5B 292.8 (65.8) 2 Ø10 (0.39) 81 (3.19) 40.07 (5.81) 2.50 (0.36) 5.62 (0.82) 81.8 (18.4) 3.58
6B 293.7 (66.0) 2 Ø10 (0.39) 81 (3.19) 40.07 (5.81) 2.50 (0.36) 5.62 (0.82) 81.8 (18.4) 3.59
3A 302.6 (68.0) 2 Ø10 (0.39) 81 (3.19) 41.5 (6.02) 2.58 (0.37) 5.80 (0.84) 84.4 (19.0) 3.59
1C 343.4 (77.2) 2 Ø10 (0.39) 81 (3.19) 52.5 (7.61) 3.15 (0.46) 7.08 (1.03) 103.0 (23.2) 3.33
2C 305.6 (68.7) 2 Ø10 (0.39) 81 (3.19) 52.5 (7.61) 3.15 (0.46) 7.08 (1.03) 103.0 (23.2) 2.97
4B 138.4 (31.1) — — 40.07 (5.81) — — — —
3C 167.7 (167.7) — — 52.5 (7.61) — — — —

Specimen b, mm (in.) heff, mm (in.) ψs,N NR,calc, kN (kip) ktest — Ratio


1B 280 (11.02) 0.9 74.87 (16.83) 3.94 — Ntest,1B/Ntest,4B 2.13

2B 280 (11.02) 0.9 74.87 (16.83) 4.06 — Ntest,2B/Ntest,4B 2.19

3B 280 (11.02) 0.9 74.87 (16.83) 4.41 — Ntest,3B/Ntest,4B 2.39

5B 380 (14.96) 0.97 109.67 (24.66) 2.67 — Ntest,5B/Ntest,4B 2.12

6B 380 (14.96) 140 0.97 109.67 (24.66) 2.68 — Ntest,6B/Ntest,4B 2.12

3A 280 (11.02) (5.51) 0.9 76.19 (17.13) 3.97 — — —

1C 280 (11.02) 0.9 85.70 (19.27) 4.01 — Ntest,1C/Ntest,3C 2.05

2C 280 (11.02) 0.9 85.70 (19.27) 3.57 — — —

4B 280 (11.02) 0.9 74.87 (16.83) 1.85 — — —

3C 280 (11.02) 0.9 85.70 (19.27) 1.96 — — —

Fig. 12—Corbel.
requirements of CEN TS 1992-4-2.4 The use of short anchor
bars allows for optimization of the manufacturing process of
the product; on the other hand, the large amount of anchor
Fig. 11—Specimens 1C, 2C, and 3C after failure.
reinforcement yielding from Eq. (1) leaves the installa-
needs to be adapted to the dimensions of the column and tion of the product relatively complicated. The anchorage
thus the product cannot be standardized. The anchorage in in Fig. 13(c) was designed based on tests presented in this
Fig. 13(b) uses a short headed bar and complies with the paper and approved by German Building Authority DIBt,11
which allows for standardized manufacturing of the product

ACI Structural Journal/January 2018 209


ratio is significantly higher in comparison to what can
be predicted by available design models.

NOTATION
d = effective depth of concrete member
ds = diameter of anchor reinforcement
fbd = design value of bond strength
fc,cube,m = mean value of concrete cube strength
fc,t,m = mean value of concrete tensile strength
heff = embedment depth of headed bar
ldt = anchorage length of headed bar
l1 = overlap length between headed bar and anchor reinforcement
NRd,a = anchorage resistance of anchor reinforcement
NRd,c = resistance against concrete cone failure
NRd,max = maximum resistance of anchor reinforcement
Ntest = maximum measured value of load
α = empirical factor (=1.0 for straight bars, =0.7 for bended bars)

AUTHOR BIOS
Ján Bujňák is a Research and Development Manager at Peikko Group
Corporation, Lahti, Finland. He received his PhD from Polytech Cler-
mont-Ferrand, Aubière, France, in 2007. His research interests include
bond, fastening technology, and punching shear in flat slabs.

Matúš Farbák is a Researcher at the University of Žilina, Žilina, Slovakia,


where he received his PhD in 2015. His research interests include the
behavior of headed bars in concrete and steel-concrete composite structures.

Fig. 13—Variants of corbel used in past and present:


REFERENCES
(a) anchorage provided with bended deformed bar; (b) 1. ACI Committee 318, “Building Code Requirements for Structural
anchorage provided by short headed bars using anchor Concrete (ACI 318-14) and Commentary (ACI 318R-14),” American
reinforcement determined in accordance with Eq. (1); and Concrete Institute, Farmington Hills, MI, 2014, 519 pp.
2. Fédération internationale du béton, “fib Bulletin 72: Bond and
(c) anchorage provided by short headed bars designed in Anchorage of Embedded Reinforcement,” Lausanne, Switzerland, 2014,
accordance with tests referenced in present paper. 161 pp.
3. EN 1992-1-1:2004, “Design of Concrete Structures—Part 1: General
and enables it to be installed into the column in a simple and Rules and Rules for Buildings,” European Committee for Standardization
productive manner. (CEN), Brussels, Belgium, 2004, 225 pp.
4. CEN/TS 1992-4-2:2009, “Design of Fastenings for Use in Concrete—
Part 4-2: Headed Fasteners,” European Committee for Standardization
CONCLUSIONS (CEN), Brussels, Belgium, 2009, 31 pp.
The presented tests demonstrated the performance of short 5. Eligehausen, R.; Malée, R.; and Silva, J. F., Anchorage in Concrete
Construction, Ernst & Sohn, Berlin, Germany, 2006, 391 pp.
headed bars with anchor reinforcement embedded in concrete 6. Henriques, J.; Raposo, J. M.; Simões da Silva, L.; and Luís Costa
columns. Such short headed bars are increasingly being used Neves, L., “Tensile Resistance of Steel-Reinforced Anchorages: Experi-
in innovative building products because they allow for opti- mental Evaluation,” ACI Structural Journal, V. 110, No. 2, Mar.-Apr. 2013,
pp. 239-249.
mization of the manufacturing process and the costs of such 7. Bujnak, J.; Farbak, M.; Bahleda, F.; and Leinonen, T., “On the Resis-
products. The main conclusions of the research are: tance of Fastening Plates with Supplementary Reinforcement,” fib Sympo-
• Short headed bars made of deformed bars with embed- sium, 2015, Copenhagen, Denmark.
8. Fromknecht, S.; Odenbreit, C.; and Dorka, U., “Versuche zur Trag-
ment depth 140 mm (5.51 in.) may anchor tensile forces fähigkeit von Ankerplatten mit einbetonierten Kopfbolzendübeln in
in a concrete column. schmalen Stahlbetonstützen,” Beton- und Stahlbetonbau, V. 105, No. 6,
• The anchorage capacity of such headed bars can be 2010, pp. 362-370. doi: (in German)10.1002/best.201000016
9. Berger, W., “Load-Displacement Behavior and Design of Anchorages
significantly increased by combining the short headed with Headed Studs with and without Supplementary Reinforcement under
bars with anchor reinforcement. Tension Load,” dissertation, University of Stuttgart, Stuttgart, Germany,
• Anchor reinforcement can be activated even with short 2015, 267 pp. (in German).
10. Bujnak, J.; Roeser, W.; Matiasko, S.; and Böhm, M., “Experimental
overlap lengths (l1 = 81 mm [3.19 in.]). Assessment of an Innovative Beam to Column Connection,” High Tech
• The ratio between the failure load of a headed bar with Concrete: Where Technology and Engineering Meet, D. Hordijk and M.
and without anchor reinforcement is 2.05 and 2.39, Luković, eds., 2018, Springer, Cham, Switzerland.
11. Deutsches Institut fur Bautechnik (DIBt), “Allgemeine bauauf-
depending on the amount of anchor reinforcement. This sichtliche Zulassung Z-21.8-2076: Peikko PCs Konsole,” 2016, Berlin,
Germany. (in German)

210 ACI Structural Journal/January 2018


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 115-S17

Evaluation of Flexural and Shear Stiffness of Concrete


Squat Walls Reinforced with Glass Fiber-Reinforced
Polymer Bars
by Ahmed Arafa, Ahmed Sabry Farghaly, and Brahim Benmokrane

Estimating the flexural and shear stiffness of concrete squat walls encountered (Paulay et al. 1982; Luna et al. 2015). Arafa et
reinforced with glass fiber-reinforced polymer (GFRP) bars is al. (2016) reported experimental results on two squat walls:
important to evaluate lateral displacement. To address this issue, one was reinforced with conventional steel bars, while
five full-scale concrete squat walls, including four reinforced with the second was reinforced with GFRP bars. The GFRP-re-
GFRP bars and one reinforced with steel bars, were tested to failure
inforced squat wall attained satisfactory strength and stable
under quasi-static reversed cyclic lateral loading. Decoupling flex-
cyclic behavior as well as self-centering capacity that
ural and shear deformations of the tested specimens showed the
contribution of shear deformation to the lateral displacement. The contributed in preventing sliding shear, which occurred in
shear stiffness of the cracked wall can be estimated based on the the steel-reinforced counterpart.
truss model with an acceptable level of conservatism. The shear- One of the most important aspects in squat-wall design
crack angle and concrete shear strength were evaluated. The is estimating wall’s lateral displacement and limiting this
flexural stiffness was estimated based on available expressions in displacement to an acceptable level. This requires an appro-
codes and guidelines related to the design of concrete members priate estimation of wall lateral stiffness, which can signifi-
reinforced with fiber-reinforced polymer bars, demonstrating their cantly affect the calculation of the natural period time and
adequacy with walls although they were established for beam and the distribution of lateral forces among structure walls as
slab elements. Based on regression analyses of the test results, well. Accordingly, estimating both the flexural and shear
expressions that correlate flexural and shear stiffness to lateral
stiffness of GFRP-reinforced squat walls was the main focus
drift were proposed. Such expressions would be vital in the context
of this study.
of displacement-based design.

Keywords: concrete squat walls; flexural and shear deformations; glass RESEARCH SIGNIFICANCE
fiber-reinforced polymer bars; seismic resistance; stiffness. This paper focuses mainly on estimating the flexural and
shear stiffness of GFRP-reinforced concrete squat walls as a
INTRODUCTION lateral seismic element. Flexural and shear deformations were
The use of fiber-reinforced polymer (FRP) materials has decoupled, showing the significant effect of shear deforma-
grown to overcome the usual problems induced by the corro- tion on the total displacement. The flexural and shear stiffness
sion of steel reinforcement in concrete structures. The inves- of the GFRP-reinforced concrete squat walls was evaluated.
tigation of FRP-reinforced concrete structures, however, Herein, these results are thoroughly discussed and compared
have focused mainly on the behavior under static-loading to the experimental results. In addition, to gain useful informa-
conditions, focusing less frequently on seismic design. tion within the context of displacement-based seismic design,
The feasibility of using FRP as internal reinforcement for expressions that directly correlate the squat-wall flexural and
lateral-resisting systems while preserving the stiffness and shear stiffness with lateral-drift ratio were proposed.
deformation capacity has become prominent. Mohamed et
al. (2014a) tested mid-rise shear walls showing the stable SUMMARY OF EXPERIMENTAL PROGRAM
cyclic performance and high level of deformability achieved AND RESULTS
by glass FRP (GFRP)-reinforced shear walls in comparison Five full-scale reinforced concrete squat walls were
to one reinforced with steel. Mohamed et al. (2014b) indi- constructed and tested to failure under quasi-static reversed
cated the potential of GFRP reinforcement in distributing cyclic lateral loading. Four specimens were entirely rein-
shear deformations along the wall height, owing to its elastic forced with GFRP bars (G4-250, G4-160, G4-80, and G6-80)
nature, resulting in control shear distortion relatively to that and one was reinforced with steel bars (S4-80). Figures 1(a)
in the steel-reinforced wall in which shear distortion took and 1(b) show the concrete dimensions and reinforcement
place simultaneously with occurrence of yielding of flex- configuration of the test specimens. The boundary elements’
ural reinforcement and mobilized at the plastic hinge zone, longitudinal and transverse reinforcement ratios and vertical
thereby deteriorating shear resistance.
ACI Structural Journal, V. 115, No. 1, January 2018.
The test results for the GFRP-reinforced mid-rise walls MS No. S-2017-052.R1, doi: 10.14359/51700987, received March 5, 2017, and
reviewed under Institute publication policies. Copyright © 2018, American Concrete
paved the way for a new experimental series using GFRP Institute. All rights reserved, including the making of copies unless permission is
bars in squat walls (shear span-to-length ratio less than obtained from the copyright proprietors. Pertinent discussion including author’s
closure, if any, will be published ten months from this journal’s date if the discussion
2.0) in which the shear-deformation problem is frequently is received within four months of the paper’s print publication.

ACI Structural Journal/January 2018 211


Fig. 1—Concrete dimensions, reinforcement details, test setup, load history, and instrumentation. (Note: 1 mm = 0.0394 in.)

Fig. 2—Failure modes of test specimens.


web reinforcement were 1.43%, 0.89%, and 0.59%, respec- sion failure (Fig. 2(b)). This was attributed to the elastic
tively, in all specimens. Four horizontal web reinforcement nature of GFRP bars, which helped the cracks to realign and
ratios equal to 0.51%, 0.79%, 1.58%, and 3.58% were used lock up in the compression zone as well as distributing shear
in G4-250, G4-160, G4-80, and G6-80 with using two layers deformations along the wall height. Figure 3 illustrates that
of No. 4 GFRP bars spaced at 250, 160, and 80 mm (9.8, both specimens exhibited similar initial stiffness. Due to
6.3, and 3.15 in.) or No. 6 GFRP bars spaced at 80 mm the relatively low elastic modulus of GFRP bars, however,
(3.15 in.), respectively. Specimen S4-80 served as a refer- G4-80 experienced softer behavior than S4-80 after initia-
ence for G4-80, so both specimens had identical reinforce- tion of the first flexural crack. The two envelopes intersected
ment configurations and ratios. One layer of bidiagonal No. at 1.35% lateral drift. Then, the strength of S4-80 deterio-
3 GFRP bars with spacing of 100 mm (3.94 in.) was added rated due to localized sliding-shear deformations, while
to prevent sliding shear. Figures 1(c) and 1(d) show the test G4-80’s strength kept increasing to achieve an ultimate load
setup and loading history, respectively. A series of linear and drift of almost 71% and 50% higher than those of S4-80,
variable differential transducers (LVDTs) and bar strain respectively.
gauges was mounted on the specimens (Fig. 1(e)). Table 1 The failure of G6-80 was identified as flexural compres-
provides the mechanical properties of the reinforcement. sion failure (Fig. 2(c)). The failure of G4-250 occurred by
Premature sliding failure dominated the behavior of S4-80 sliding along a major diagonal shear crack (Fig. 2(d)) due
due to flexural reinforcement yielding, which produced a to the inadequacy of horizontal web reinforcement, while
major horizontal crack (Fig. 2(a)). Such behavior, however, G4-160 experienced sudden flexural rupture in the longitu-
was prevented in G4-80, which exhibited flexural compres- dinal bars at the boundary element under tension (Fig. 2(e)).

212 ACI Structural Journal/January 2018


Table 1—Tensile properties of reinforcement
Tensile modulus of
Bar Designated bar diameter, mm Nominal area*, mm2 elasticity, GPa Tensile strength†‡, MPa Average strain at ultimate, %
Straight bars
No. 3 GFRP 9.5 71 65 1372 2.1
No. 3 steel 9.5 71 200 fy = 420 εy = 0.2
No. 4 steel 12.7 129 200 fy = 420 εy = 0.2
Bent No. 3 GFRP—rectilinear spiral
Straight 50 1065 2.1
9.5 71
Bent — 460 —
Bent No. 4 GFRP—horizontal bar
Straight 50 1020 2.0
12.7 129
Bent — 459 —
Bent No. 6 GFRP—horizontal bar
Straight 50 1028 2.0
19.1 285
Bent — 463 —
*
According to CSA S807 (CSA 2010).

Tensile properties were calculated using nominal cross-sectional areas.

Guaranteed tensile strength: average value is three times standard deviation (ACI 440.1R-15).
Notes: 1 mm = 0.0394 in.; 1 MPa = 145 psi; fy is steel yielding strength; εy is steel yielding strain.

in the length of the two diagonals using Eq. (1a). The flex-
ural deformations, on the other hand, can be calculated based
on the two vertical LVDTs mounted at both boundaries with
height h using Eq. (1b) (Fig. 4(a))

(d1′ − d )d − (d 2′ − d )d
U s original = γ original h = (1a)
2hL

(VL − VR )
Uf original = θh = h (1b)
L

where γoriginal is the shear distortion over a height h; d1′ and


d2′ are the lengths of the diagonal LVDTs after deformation;
d is the original length of the diagonal before deformation;
L and h are the length and height of the panel, respectively;
Fig. 3—Load-top displacement envelope curves. (Note: 1 mm and θ is the rotation over the height h.
= 0.0394 in.; 1 kN = 0.225 kip.) The deformation calculated by this method, however, is
Figure 3 shows that the horizontal web reinforcement ratio overestimated because of the variation in bending moment
had a significant effect on increasing the ultimate strength along the wall height. Hiraishi (1984) demonstrated that this
and drift ratio. This effect, however, appears to have been method is only valid if the center of rotation is located at the
insignificant when the wall was provided with more hori- element center (Fig. 4(b)), which is not the case of structural
zontal web reinforcement than required for flexural resis- walls. Therefore, a portion of the change in diagonal lengths
tance (G6-80 compared to G4-80). Overall, the observed has to be attributed to flexural deformations (Fig. 4(c)). More-
behavior reveals the acceptable behavior of GFRP-rein- over, Eq. (1b) is only valid if the curvature is concentrated at
forced walls as a lateral-resisting system in low to moderate the wall base, which is not the case in many walls. Therefore,
earthquake regions. Table 2 summarizes the failure progres- Hiraishi (1984) suggested the following corrected equations
sion of the test specimens.
Us corrected = Us original – (α – 0.5)θh (2a)
DECOUPLING OF FLEXURAL AND SHEAR
DEFORMATIONS Uf corrected = αθh = αUf original (2b)
Oesterle et al. (1979) suggested that the shear deformation
∫0 θ ( y ) dy
h
of a shear panel can be directly estimated from the changes α= (2c)
θh

ACI Structural Journal/January 2018 213


Table 2—Summary of test results
Initial flexural Initial shear Concrete cover Excessive cover
crack crack Steel yielding splitting spalling Peak capacity Failure
Wall fc′,
designation MPa P, kN d, % P, kN d, % P, kN d, % P, kN d, % P, kN d, % P, kN d, % P, kN d, %
G4-250 35 147 0.025 228 0.39 — — 510 1.6 615 2.0 678 2.65 678 2.65
G4-160 35 145 0.025 227 0.36 — — 528 1.5 639 2.0 708 2.8 708 2.8
G4-80 40 164 0.026 234 0.34 — — 567 1.5 710 2.0 912 2.75 810 3.0
G6-80 41 168 0.026 237 0.29 — — 582 1.5 739 2.0 935 2.9 815 3.1
S4-80 35 160 0.025 230 0.1 425 0.4 525 1.0 534 1.25 534 1.25 440 2.00

Notes: fc′ is concrete compressive strength; P is applied lateral load; d is drift ratio; 1 MPa = 145 psi; 1 kN = 0.225 kip.

Fig. 4—Decoupling of flexural and shear deformations.


Original and corrected flexural and shear
deformations
As shown in Fig. 7, the total displacement was substan-
tially overestimated with Eq. (1), while the results with
Eq. (2) were in good agreement with the measured lateral
displacement. Figure 8 shows plots of the corrected flex-
ural and shear deformations. The behavior of all specimens
was dominated initially by flexural response. With the initi-
ation of the first shear crack, shear deformations began to
contribute in the total displacement. Apparently, the contri-
bution of shear deformation varied as a function of reinforce-
ment type (steel or GFRP) and horizontal web reinforcement
ratio. In S4-80, the percentage of shear deformation to the
Fig. 5—Method for estimating α (rotation profile over wall total deformation corresponding to yielding of the longitu-
height). dinal reinforcement was 22% (0.4% drift) and increased to
46% after the localization of sliding shear corresponding
The center of rotation (α) accounts for the variation in curva- to concrete cover spalling (1.25% drift) and reached 64%
ture along the panel height and can be estimated as shown at failure (2% drift). In the GFRP-reinforced specimens,
in Fig. 5. Mohamed et al. (2014b) stated that Eq. (1) over- the percentages were approximately 49%, 38%, 28%, and
estimated shear deformations, while correcting the results 15% for G4-250, G4-160, G4-80, and G6-80, respectively,
based on Eq. (2) produced consistent results. This confirms up to 1% drift. At cover spalling (2% drift), the percentage
the necessity of involving α in the calculation to obtain increased slightly to 56%, 42%, 36%, and 20%, respectively,
corrected values for decoupled deformations. In this study, which remained almost constant up to failure. The results
the flexural and shear deformations were calculated over a reveal that shear deformation should not be neglected, even
height h equal to the wall length lw based on the arranged if the shear strength was twice the applied load, as is the case
instrumentation (Fig. 1(e)). To assist in calculating the with G6-80.
center of rotation, average curvature and rotations over the All specimens exhibited negligible initial flexural deforma-
wall height were calculated (Fig. 6). α was found to be 0.62 tion because the gross flexural stiffness resisted the deforma-
for S4-80, 0.59 for G6-80, and 0.58 for the other specimens. tion, as shown in Fig. 9(a). By the onset of the first flexural

214 ACI Structural Journal/January 2018


Fig. 6—Calculated curvature and rotation profiles for test specimens. (Note: 1 mm = 0.0394 in.)

Fig. 7—Measured and calculated displacement at height equal to wall length. (Note: 1 mm = 0.0394 in.)

Fig. 8—Displacement components at height equal to wall length. (Note: 1 mm = 0.0394 in.)
crack, the specimens experienced a reduction in lateral stiff- Figures 9(c) and 9(d) show that, similar to the flexural-de-
ness, resulting in a significant increase in flexural deforma- formation response, initial high shear stiffness was followed
tion. The flexural deformation in all GFRP-reinforced speci- by a significant reduction in shear stiffness manifested with
mens followed a similar trend, showing no significant effect of the appearance of the first shear crack. The figures under-
changing the horizontal web reinforcement ratio. Figure 9(b) line the link between steel yielding and shear deformations
shows the normalized secant flexural stiffness to the initial in S4-80, which exhibited significant degradation in shear
elastic flexural stiffness. Due to the relatively low modulus stiffness after a few cycles of steel yielding associated with
of elasticity of GFRP bars, the stiffness loss in the GFRP-re- substantial increasing in shear deformation. In contrast, the
inforced squat walls was relatively pronounced compared to shear deformation in GFRP-reinforced squat walls increased
the steel-reinforced one. It is worth mentioning, however, that almost linearly with loading. The figures also reveal the
the softer behavior of GFRP-reinforced walls will increase the effectiveness of the horizontal web reinforcement ratio in
displacement demand, which could be considered an advan- reducing shear deformation.
tage because a softer structure attracts lower seismic forces.

ACI Structural Journal/January 2018 215


Fig. 10—Truss model for shear-deformation estimation.
where Δs is the shear deformation at height h; θ is the angle
between the shear crack and the longitudinal axis of the
wall; n is the modular ratio = Es/Ec; Es and Ec are the elastic
Fig. 9—Comparison between flexural and shear deforma- moduli of the horizontal web reinforcement and concrete,
tions in test specimens. respectively; ρh is the horizontal web reinforcement ratio; Vs
is the shear force carried by horizontal web reinforcement;
SHEAR STIFFNESS b is the wall thickness; d is the effective shear depth, which
Although predicting the shear stiffness of test specimens could be taken as 0.8 of the wall length; fs is the stress in
beyond shear cracking is necessary, the FRP design codes and horizontal web reinforcement; and Ah is the area of hori-
guidelines (CSA S806 [2012], ACI Committee 440 [2015]) zontal web reinforcement within a spacing s.
contain no seismic provisions. Therefore, the methods used The shear stiffness is calculated as the ratio between the
in estimating the cracked shear stiffness of steel-reinforced applied load (V) and Δs, as follows
squat walls were reviewed and the most appropriate method
to apply to GFRP-reinforced squat walls was selected.
V ρ sin 2 θ ⋅ cos 2 θ b ⋅ d ⋅ Es ⋅ V
The shear stiffness in S4-80 significantly decreased after Ks = = h 4 (5)
the first shear crack and reduced to 58% at the onset of flex- ∆s sin θ + nρh Vs ⋅ h
ural-reinforcement yielding, followed by severe deteriora-
tion due to the development of sliding-shear deformations The model calculates the shear stiffness of steel-rein-
(Fig.  9(d)), confirming past research outcomes (Salonikios forced walls based on shear-reinforcement yielding and a
et al. 1999, Massone et al. 2009). According to Eurocode shear crack angle (θ) conservatively assumed to be equal
8 (2004) and ASCE/SEI 43-05 (2005), the cracked shear to 45 degrees. The model cannot be directly applied to
stiffness is taken as 50% of the elastic shear stiffness, while GFRP-reinforced squat walls due to the absence of a yielding
ACI 318 (2014) and CSA A23.3 (2014) account only for point and the significant deviation of θ from 45 degrees in
the reduction in flexural stiffness, allowing no reduction many test series. The current concept for calculating shear
in the cracked shear stiffness. Given this disagreement and resistance in most guidelines and codes is to find a rational
uncertainty, Li and Xiang (2011) proposed and experimen- estimation for the shear carried by the concrete Vc; then, the
tally verified an analytical model based on truss analogy shear-reinforcement (Vs) portion can be calculated. ACI 318
to predict the cracked shear stiffness corresponding to the (2014) assumes that Vc is equal to the first shear-cracking
yielding point (Fig. 10). Tang and Su (2014) undertook load (Vcr). On the other hand, ACI 440.1R-15 (2015), CSA
further verification of the model. S6-14 (2014), and CSA S806-12 (2012) provide formulas
The Li and Xiang (2011) model assumes that shear defor- for calculating Vc and θ.
mation is the summation of tie elongation (horizontal web
reinforcement) and shortening of concrete struts, which Design codes and guides for shear-strength
could be formulated as follows predictions
ACI guide provision (ACI 440.1R-15)—The estimation of
sin 4 θ + nρh Vs ⋅ h Vc in ACI 440.1R-15 for FRP-reinforced concrete elements
∆s = 2 2
(3) is based on the assumption that θ is 45 degrees and that Vc
ρh sin θ ⋅ cos θ b ⋅ d ⋅ Es
can be calculated as follows
f s ⋅ dcotθ ⋅ Ah
Vs = (4) Vc = 2/5√fc′bwc; c = kd, k = 2 ρ f n f + (ρ f n f ) 2 − ρ f n f ,
s
and n = Ef/Ec (6)

216 ACI Structural Journal/January 2018


Fig. 11—Effect of shear-crack angle.
where bw is the wall thickness; c is the neutral-axis depth It should be noted that θ in Eq. (12) is limited to 50 degrees
of the cracked transformed section; and d is the effective based on the limitation of εx in Eq. (11).
flexural depth.
CSA code provisions (S806-12 and S6-14)—According to Evaluation of shear-crack angle (θ)
CSA S806 (2012), the Vc of flexural members can be evalu- ACI 440.1R-15 assumes θ to be 45 degrees, while the
ated as follows calculation of θ based on CSA S6-14 and CSA S806-12
resulted in 50 and 60 degrees, respectively. A unified method
Vc = 0.05 km kr ka k s 3 f c′ bw d v (7) for calculating Vs was used (ACI Committee 318 2014) in
which the concrete shear strength (Vc) was set to be equal to
where the experimental cracking load (Vcr) (given in Table 2). The
shear deformation prior to cracking was calculated using the
0.11 f c′bw d v ≤ Vc ≤ 0.22 f c′bw d v ; km = V f d /M f ≤ 1.0; elastic shear stiffness defined by Park and Paulay (1975) for
prismatic elements as follows
kr = 1 + 3 E f ρ f ; ka = 2.5/(Mf/Vfd) ≥ 1.0;
Gc Ash
ks = 750/(450 + d) ≤ 1.0 (8) K se = (13)
f ⋅ hw
where dv is the effective shear depth, taken as the greater
of 0.9d or 0.72lw and not less than 0.8lw according to CSA where Gc is the concrete shear modulus (= 0.4Ec); Ash is the
A23.3-14; and Mf and Vf are the moment and shear applied effective shear area of an uncracked element (=(5 + 5ν)/(6 +
at the critical section of shear, respectively. 5ν)Ag), where Ag is the gross-section area and ν is Poisson’s
In contrast to ACI 440.1R-15, CSA S806-12 provides ratio; and f accounts for nonuniform distribution of the shear
Eq. (9) to estimate θ stresses and is equal to 1.2 for rectangular cross sections.
Figure 11 shows that adopting a shear-crack angle of 45
30 degrees ≤ θ = 30 + 7000εl ≤ 60 degrees; and 50 degrees as proposed in ACI 440.1R-15 and CSA
M f /d v + V f S6-14 underestimated the shear deformations by 53% and
εl = (9) 35%, respectively, because the experimentally measured
2 E f Af
angles ranged from 55 to 59 degrees. Thus, θ < 50 degrees
could lead to unsafe predictions. Increasing this limitation to
According to CSA S6-14, the Vc of flexural members using 60 degrees, as proposed in CSA S806-12, however, yielded
FRP as the main reinforcement can be evaluated as follows higher shear displacement than the experimental results with
a mean value of 36%. This revealed that 56 degrees was the
 0.4 1300  angle that produced the best fit with the experimental results
Vc =  × φc f c′ bw d v (10)
 1 + 1500 ε x 1000 + sxe  (Fig. 11). However, this angle might lead to unsafe predic-
tions in other test series. Therefore, the shear-crack angle
based on CSA S806-12 (θ = 60 degrees) is conservatively
M f /d v + V f 35 d v
where ε x = ≤ 0.003, sxe = ≥ 0.85 d v , recommended in this study.
2 E f Af 15 + ag
and ag is the aggregate size (11) Evaluation of concrete shear strength (Vc)
Vc was predicted based on ACI 440.1R-15, CSA S6-14,
CSA S6-14 provides an equation similar to that in CSA and CSA S806-12 as well as Vc = Vcr while setting θ to 60
S806-12 to calculate θ as follows degrees. Figure 12 shows that the predictions according to
ACI 440.1R-15 and CSA S6-14 overestimated shear defor-
θ = (29 + 7000εx)(0.88 + Sze/2500) (12) mations with mean values of 74% and 64%, respectively.
While the predicted Vc based on CSA S806-12 resulted in
closer agreement with the mean value of 13%, the prediction

ACI Structural Journal/January 2018 217


Fig. 12—Effect of concrete shear strength (at θ = 60 degrees). (Note: 1 mm = 0.0394 in.; 1 kN = 0.225 kip.)
Proposed shear-stiffness model
Seismic design practice worldwide is moving toward
displacement-based design methods. Therefore, the seismic
demands of a structural system should be evaluated using an
effective natural period of a structure based on the effective
secant stiffness connected to a targeted displacement point
on the load-displacement envelope curve. Hence, it would
be preferable, within the context of a displacement-based
design method, to develop a simple model that directly
correlates the shear-stiffness degradation of a wall to its
top drift. Accordingly, the authors predicted the normalized
shear-stiffness degradation based on regression analysis.
The results in Fig. 15(a) generally reflect strong correla-
tion between normalized shear stiffness and drift ratio
(R2 = 0.87). Interpretation of the results also revealed that
normalized shear-stiffness degradation is a function of hori-
Fig. 13—Concrete shear strength versus top displacement. zontal web reinforcement ratio. As shown in Fig. 15(b), the
(Note: 1 mm = 0.0394 in.; 1 kN = 0.225 kip.) degradation decreased almost linearly with increasing rein-
forcement ratio at the same drift level. The relation appears,
may lead to unsafe estimation at elastic stage prior to shear however, to have significantly deviated with increasing drift.
cracking load, which is significantly lower than the calcu- This is clear in Fig. 15(c) and (d), which show the nonlinear
lated shear strength and resulted in underestimated shear variation of relation constants A and B with the drift ratio,
deformation (Fig. 12). indicating the interrelation between the drift ratio and hori-
The experimental concrete shear strength Vc was calcu- zontal web reinforcement ratio.
lated as Vc = V − VFRP, where V is the applied load and VFRP Based on this discussion, the normalized shear-stiffness
is the contribution of the FRP horizontal web reinforce- degradation was set as a function of top drift ratio and hori-
ment calculated using the truss analogy (Eq. (4)) with the zontal web reinforcement ratio as follows
average measured shear-strain values. Although using a
constant value for Vc is not representative in estimating shear Ks
deformation as Vc varies with loading (Fig. 13), it could be = a ⋅ δ x ⋅ ρ + b ⋅ δ y (14)
K se
conservatively assumed that Vc = Vcr is equal to the first shear
cracking load. Therefore, it would be appropriate to deter-
mine a formula to estimate Vcr. In this regard, the ACI 318 where Ks is the secant shear stiffness at a lateral top drift
(2014) procedure for concrete shear strength was followed equal to δ; Kse is the elastic shear stiffness and can be calcu-
in which the cracking load is normalized to bwlwfc′0.5. Based lated with Eq. (13); ρ is the horizontal web reinforcement
on the experimental shear-cracking-load values (Table 2), ratio; and a, b, x, and y are constants. For the experimental
the concrete shear strength (Vc) could be estimated as an data, the coefficients a, b, x, and y, which correlate between
average value equal to 0.12bwlwfc′0.5. experimental and analytical results, were found to be equal to
Figure 14 confirms the underestimation using the equation 0.04, 0.03, –1, and –1.6, respectively. Therefore, the normal-
in ACI 440.1R-15 and CSA S6-14, while the CSA S806-12 ized shear stiffness (Ks/Kse) can be rewritten as follows
equation initially overestimated the shear stiffness degrada-
tion and got noticeably better predictions at high load levels K s 0.04ρ 0.03
= + 1.6 (15)
exceeding Vcr. Furthermore, the calculation based on Vc = K se δ δ
0.12bwlwfc′0.5 produced better estimations with an acceptable
level of conservatism. Figure 16 shows good agreement between the exper-
imental results and the proposed formula. More testing is
needed, however, to clarify the effect of other parameters

218 ACI Structural Journal/January 2018


Fig. 15—Validation of proposed model for shear-stiffness
Fig. 14—Shear-stiffness degradation normalized with gross degradation.
shear stiffness versus drift ratio.
such as aspect ratio, axial loading, and the presence of
boundary elements on the normalized shear stiffness.

FLEXURAL STIFFNESS
In practice, code methods employ an effective flexural
stiffness to consider the reduction in stiffness caused by
cracking for reinforced concrete elements. In designing
steel-reinforced walls, a single reduction factor (α) is
commonly applied to the gross stiffness to reduce it to effec-
tive stiffness. For instance, the effective flexural stiffness for
a wall under zero axial load is proposed to be 25%, 35%, and
50% of the gross flexural stiffness by NZS 3101 (1995), ACI
318-14, and Eurocode 8 (2004), respectively. CSA A23.3-14
links α to the ductility- and overstrength-related force modi-
fication factors (Rd and Ro) as follows

α = 1 – 0.35(RdRo/γw – 1), 0.5 ≤ α ≤ 1 (16)

where γw is the wall overstrength factor and can be taken


equal to Ro. According to CSA A23.3-14 and NBCC (2010),
Rd and Ro for moderately ductile steel-reinforced squat walls Fig. 16—Comparison of predicted normalized shear stiff-
are 2 and 1.4, respectively. Substituting these values in ness (Eq. (5)) with experimental data.
Eq. (16) results in α = 0.65.
the same reduction factors to estimate the effective flex-
As mentioned earlier, the measured secant flexural stiff-
ural stiffness of the GFRP-reinforced squat walls. This was
ness for specimen S4-80 was almost 30% of the gross flex-
expected and can be attributed to the different mechanical
ural stiffness corresponding to steel yielding. Comparing the
and bond characteristics between steel and GFRP bars.
value at yielding with the aforementioned suggestions for
Instead, the methods for estimation of flexural stiffness for
stiffness-reduction values reveals that Eurocode 8 (2005)
FRP-reinforced elements specified in ACI 440.1R-15 and
and CSA A23.3 (2014) overestimated α by 67% and 116%,
CSA S806-12 were employed.
respectively, while ACI 318 (2014) and NZS 3101 (1995)
ACI 440.1R-15 provides an expression for estimating the
yielded better estimations, with differences of 17% higher
effective moment of inertia (Ie) to calculate the deflection of
and lower than the measured stiffness, respectively. Based
FRP-reinforced concrete elements; the formula is given as
on Fig. 9(b), however, it would appear inappropriate to use
follows

ACI Structural Journal/January 2018 219


Fig. 17—Predicted flexural displacement and normalized flexural stiffness versus experimental data.

I cr
Ie = ≤ I g (17)
1 − γ ( M cr /M a ) 2 [1 − I cr /I g ]

where Icr is the cracked moment of inertia; γ is a function of


the cracking moment (Mcr) relative to the applied moment
(Ma) and equal to [3 – 2(Mcr/Ma)] for a cantilever with
end-point load; and Ig is the gross moment of inertia.
The method proposed in CSA S806-12 is based on inte-
grating the moment-curvature relation, which is assumed to
stay linear under increased loading with a flexural rigidity
of EcIg in the uncracked stage and EcIcr in the cracked stage,
whereas there is no tension stiffening in the cracked zone
of a beam element. The effective moment of inertia can be
calculated as follows
Fig. 18—Normalized flexural stiffness versus drift ratio.

I cr together in Fig. 17 along with the results obtained experimen-


Ie = ≤ I g (18)
1 − η( Lg /L)3 tally. Using the calculated flexural deformations, the secant
flexural stiffness at different drift levels was also calculated
where η equals (1 – Icr/Ig); and Lg is the distance from the and normalized to the gross flexural stiffness and plotted
free end to the point at which M = Mcr. in Fig. 17. It should be noted that calculations were made
Knowing the effective moment of inertia (Eq. (17) and using the experimentally obtained cracking moment corre-
(18)), the flexural deformations at any load level can be sponding to the first flexural crack, as given in Table 2. It can
calculated using the elastic deflection equation as follows be inferred from the figure that using either ACI 440.1R-15
or CSA S806-12 in calculating flexural deformation and the
P ⋅ x2 consequent normalized flexural stiffness yielded good agree-
δ= (3l − x) (19)
6 EI e ment with the experimental results, although CSA S806-12
appears to be more conservative.
where P is the applied load; δ is the lateral displacement; E Similar to shear stiffness and based on regression analysis,
is the elastic modulus for concrete, which can be calculated the normalized flexural stiffness could be formulated as in
as 4700√fc′ MPa (57,000√fc′ psi); x is the distance from the Eq. (20). Figure 18 verifies the proposed equation
support to the point at which the displacement is calculated
and equal to 1500 mm (59 in.); and l is the shear span equal Kf/Kfe = 0.1δ–0.6 (20)
to 2550 mm (100.4 in.).
The predicted flexural deformations using the effective CONCLUSIONS
moment of inertia specified by ACI 440.1R-15 and CSA The main purpose of this research was to evaluate the
S806-12 were calculated at different load levels and plotted flexural and shear stiffness of GFRP-reinforced concrete

220 ACI Structural Journal/January 2018


squat walls. The evaluation process and its outcomes can be REFERENCES
concluded as follows: ACI Committee 318, 2014, “Building Code Requirements for Struc-
tural Concrete (ACI 318-14) and Commentary (ACI 318R-14),” American
1. Decoupling flexural and shear deformation without Concrete Institute, Farmington Hills, MI, 519 pp.
considering curvature variation can lead to overestimated ACI Committee 440, 2015, “Guide for the Design and Construction of
deformation, while correcting the deformations based on the Concrete Reinforced with FRP Bars (ACI 440.1R-15),” American Concrete
Institute, Farmington Hills, MI, 44 pp.
estimated center of rotation produced consistent results. Arafa, A.; Farghaly, A. S.; and Benmokrane, B., 2016, “Experimental
2. Shear deformation contributed significantly to the total Investigation of GFRP-Reinforced Squat Wall Subjected to Lateral Load,”
deformation and cannot be omitted, even if the shear strength 7th International Conference on Advanced Composite Materials in Bridges
and Structures (ACMBS-VII), Vancouver, BC, Canada, 10 pp.
is greater than the applied load. ASCE/SEI 43, 2005, “Seismic Design Criteria for Structures, Systems,
3. The cracked shear stiffness of GFRP-reinforced and Components in Nuclear Facilities (ASCE/SEI 43-05),” American
concrete squat walls can be estimated based on the truss Society of Civil Engineers, Reston, VA.
CSA A23.3-14, 2014, “Design of Concrete Structures Standard,” Cana-
model with an acceptable level of conservatism. To ensure dian Standards Association, Mississauga, ON, Canada, 240 pp.
this conservatism, the shear-crack angle should be estimated CSA S6-14, 2014, “Canadian Highway Bridge Design Code,” Canadian
based on CSA S806-12 and the concrete shear strength is Standards Association, Mississauga, ON, Canada, 732 pp.
CSA S806-12, 2012, “Design and Construction of Building Components
recommended to be equal to 0.12bwlwfc′0.5. with Fiber-Reinforced Polymers,” Canadian Standards Association, Missis-
4. In calculating flexural deformation and the consequent sauga, ON, Canada, 208 pp.
normalized flexural stiffness, both ACI 440.1R-15 and CSA Eurocode 8, 2004, “Design of Structures for Earthquake Resistance, Part
1: General Rules, Seismic Actions and Rules for Buildings, ENV 1998-
S806-12 yielded results in good agreement with the exper- 1:2003,” Comité Européen de Normalisation, Brussels, Belgium.
imental results, although CSA S806-12 appears to be more Hiraishi, H., 1984, “Evaluation of Shear and Flexural Deformations of
conservative. Flexural Type Shear Walls,” Bulletin of the New Zealand National Society
for Earthquake Engineering, V. 17, No. 2, pp. 135-144.
5. Within the context of displacement-based design method, Li, B., and Xiang, W., 2011, “Effective Stiffness of Squat Structural
a simple model that correlates the flexural and shear stiffness Walls,” Journal of Structural Engineering, ASCE, V. 137, No. 12, pp. 1470-
degradation of the test walls to their drift was proposed. 1479. doi: 10.1061/(ASCE)ST.1943-541X.0000386
Luna, B. N.; Rivera, J. P.; and Whittaker, A. S., 2015, “Seismic Behavior
of Low-Aspect-Ratio Reinforced Concrete Shear Walls,” ACI Structural
AUTHOR BIOS Journal, V. 112, No. 5, Sept.-Oct., pp. 593-604. doi: 10.14359/51687709
Ahmed Arafa is a Doctoral Candidate in the Department of Civil Engi- Massone, L. M.; Orakcal, K.; and Wallace, J. W., 2009, “Modeling
neering at the University of Sherbrooke, Sherbrooke, QC, Canada. He of Squat Structural Walls Controlled by Shear,” ACI Structural Journal,
received his BSc and MSc degrees from Assiut University, Assiut, Egypt. V. 106, No. 5, Sept.-Oct., pp. 646-655.
His research interests include testing of concrete structures reinforced with Mohamed, N.; Farghaly, A. S.; Benmokrane, B.; and Neale, K. W.,
fiber-reinforced polymers. 2014a, “Experimental Investigation of Concrete Shear Walls Reinforced
with Glass-Fiber-Reinforced Bars under Lateral Cyclic Loading,” Journal
Ahmed Sabry Farghaly is a Postdoctoral Fellow in the Department of of Composites for Construction, ASCE, V. 18, No. 3.
Civil Engineering at the University of Sherbrooke. His research interests Mohamed, N.; Farghaly, A. S.; Benmokrane, B.; and Neale, K. W.,
include nonlinear analysis of reinforced concrete structures and behavior of 2014b, “Flexure and Shear Deformation of GFRP-Reinforced Shear Walls,”
structural concrete reinforced with fiber-reinforced polymers. Journal of Composites for Construction, ASCE, V. 18, No. 2.
NBCC, 2010, “National Building Code of Canada,” Canadian Commis-
Brahim Benmokrane, FACI, is Professor of civil engineering and NSERC sion on Building and Fire Codes, National Research Council of Canada,
Research Chair in FRP Reinforcement for Concrete Infrastructure and Montreal, QC, Canada
Tier-1 Canada Research Chair in Advanced Composite Materials for Civil NZS 3101, 1995, “Code of Practice for the Design of Concrete Struc-
Structures in the Department of Civil Engineering at the University of Sher- tures, Part 1,” Standards New Zealand, Wellington, New Zealand.
brooke. He is a member of ACI Committee 440, Fiber-Reinforced Polymer Oesterle, R. G.; Aristizabal-Ochoa, J. D.; Fiorato, A. E.; Russell, H. G.;
Reinforcement. and Corley, W. G., 1979, “Earthquake Resistant Structural Walls—Test of
Isolated Walls—Phase II.” Report ENV77-15333, National Science Foun-
dation, Arlington, VA.
ACKNOWLEDGMENTS Park, R., and Paulay, T., 1975, Reinforced Concrete Structures, John
The experimental study was conducted with funding from the Tier-1 Wiley & Sons, Inc., New York.
Canada Research Chair in Advanced Composite Materials for Civil Struc- Paulay, T.; Priestley, M. J. N.; and Synge, A. J., 1982, “Ductility in
tures, and Natural Sciences and Engineering Research Council of Canada Earthquake Resisting Squat Shear Walls,” ACI Journal Proceedings, V. 79,
(NSERC-Industry Research Chair program), and the Fonds de recherche du No. 4, Apr., pp. 257-269.
Québec – Nature et Technologie (FRQ-NT). The assistance of the technical Salonikios, T. N.; Kappos, A. J.; Tegos, I. A.; and Penelis, G. G., 1999,
staff of the new Canadian Foundation for Innovation (CFI) structural lab “Cyclic Load Behavior of Low-Slenderness Reinforced Concrete Walls:
at the University of Sherbrooke’s Department of Civil Engineering is also Design Basis and Test Results,” ACI Structural Journal, V. 96, No. 4,
acknowledged. July-Aug., pp. 649-660.
Tang, T. O., and Su, R. K. L., 2014, “Shear and Flexural Stiffnesses of
Reinforced Concrete Shear Walls Subjected to Cyclic Loading,” The Open
Construction and Building Technology Journal, V. 8, pp. 104-121. doi:
10.2174/1874836801408010104

ACI Structural Journal/January 2018 221


ARE YOU A RESEARCHER?
SIGN UP FOR TODAY!
ORCID provides a persistent digital identifer that distinguishes you
from every other researcher and, through integration in key research
workflows such as manuscript and grant submission, supports
automated linkages between you and your professional activities,
ensuring that your work is recognized.
Individuals may use ORCID services freely and it’s as easy as 1-2-3:

1 REGISTER

2 ADD YOUR INFO

3 USE YOUR ORCID ID

For more information and to register, visit:


WWW.ORCID.ORG
ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 115-S18

Long-Term Multipliers and Deformability of Fiber-


Reinforced Polymer Prestressed Concrete
by Yail J. Kim and Raymon W. Nickle

This paper presents the flexural characteristics of highway bridge characteristics of vertically distributed CFRP tendons, when
girders prestressed with fiber-reinforced polymer (FRP) tendons. the prestressed concrete beams were subjected to bending.
Of interest are the technical challenges identified by ACI Subcom- Unlike steel strands, the stress development of CFRP was
mittee 440-I (FRP-Prestressed Concrete): long-term multipliers, substantially influenced by the degree of curvature that led
deformability, and minimum reinforcement. Aramid and carbon
to the progressive rupture of the distributed tendons. Saiedi
FRP (AFRP and CFRP, respectively) composites are used to
et al.8 evaluated the performance of CFRP-prestressed
prestress concrete girders. Based on analytical models, new design
expressions are proposed for the aforementioned items, followed by concrete beams in an aggressive environment. Fatigue
an assessment using laboratory test data and full-scale benchmark loading at a low temperature of –28°C (–18°F) degraded the
bridges. The long-term multipliers calibrated per reliability theory beams’ stiffness and camber, which accompanied the slip
are mostly different from the empirical multipliers adopted in of the CFRP. Zhang et al.9 reported the axial response of
ACI 440.4R-04. The girders prestressed with AFRP/CFRP suffi- multiple CFRP tendons for prestressing concrete members.
ciently deform in flexure, even though their moment-curvature Parameters affecting the CFRP behavior were tendon stiff-
responses are not comparable with those of steel-prestressed ness, anchoring methods, and geometric properties. Further
girders. The new deformability index specifies design require- details about FRP-prestressed members, such as bond and
ments for AFRP/CFRP-prestressed members with either compres- transfer length, anchorage, durability, and field applications
sion-controlled or tensioned-controlled sections. The importance
are available in review papers.10,11
of a potential change in FRP modulus during the service life of
Concerning long-term deflections for FRP-prestressed
prestressed concrete girders is examined. A factored ultimate-to-
cracking moment ratio of ϕMn/Mcr = 1.2 is suggested for girders concrete, empirical multipliers are exploited in current
prestressed with AFRP/CFRP, which aligns with the articles of practice.3 The origin of these multipliers can be found in an
existing design manuals and specifications dedicated to prestressed analysis using steel-prestressed concrete, reported over 40
concrete. years ago.12 Although the developer clarified the inherent
unpredictability of long-term deflections, the multipliers are
Keywords: bridge; deformability; fiber-reinforced polymer (FRP); long- adopted in most prestressed concrete applications, including
term multiplier; minimum reinforcement; prestress.
those with FRP tendons, owing to the absence of relevant
research.
INTRODUCTION The deformability of FRP-prestressed members has
To increase the durability of prestressed concrete bridges, been studied by a few researchers. It is important to note
it has been proposed to use advanced composites instead that the concept of ductility in traditional concrete struc-
of conventional steel strands.1,2 Fiber-reinforced polymers tures is not valid for FRP-prestressed members due to
(FRP), comprising synthetic fibers and a resin matrix, are FRPs’ nonyielding nature; accordingly, an alternative term
promising materials with numerous advantages—namely, called “deformability” is employed. Zou13 tested concrete
low density at high strength, noncorrosiveness, and noncon- beams prestressed with AFRP and CFRP tendons, and
ductance.3 As prestressing tendons, aramid FRP (AFRP) compared their behavior in terms of curvature development,
and carbon FRP (CFRP) are used in preference. Glass sectional rotation, and displacement. A deformability factor
FRP (GFRP) is inadequate for prestressing applications was proposed by combining displacement and moment at
because of unfavorable creep resistance.4 Nanni and Tani- cracking and ultimate stages of the prestressed beams. The
gaki5 showed one of the early endeavors in FRP-prestressed applicability of this factor, however, appears to be limited
concrete. After testing 21 beams with several variables because rigorous computation is necessary to determine
(tendon size, prestress level, and shear reinforcement), it was the beams’ deflections. Au and Du14 developed a numerical
concluded that FRP was usable for prestressing concrete; model to predict the flexural behavior of FRP-prestressed
however, further development was recommended. Grace and beams. A comparative study was performed to assess the
Abdel-Sayed6 proposed a prestressing concept for highway variation trend of existing deformability expressions with
bridges with a combination of bonded and unbonded CFRP
tendons. A reduced-scale prototype bridge superstructure
ACI Structural Journal, V. 115, No. 1, January 2018.
was constructed and tested in static, dynamic, and cyclic MS No. S-2017-054, doi: 10.14359/51700988, received February 20, 2017, and
loadings. The performance of CFRP was satisfactory; for reviewed under Institute publication policies. Copyright © 2018, American Concrete
Institute. All rights reserved, including the making of copies unless permission is
example, no visual damage was noticed after 7 million obtained from the copyright proprietors. Pertinent discussion including author’s
closure, if any, will be published ten months from this journal’s date if the discussion
fatigue cycles. Dolan and Swanson7 examined the failure is received within four months of the paper’s print publication.

ACI Structural Journal/January 2018 223


reinforcement ratio. Because the expressions were inde- 625 and 1460 MPa (90 and 212 ksi), respectively, in compli-
pendently developed, meaningful conclusions were not ance with ACI 440.4R-04.3
accomplished. ACI 440.4R-043 specifies a deformability
equation, whereas its applicability is practically restricted Simulation approach
due to the facts that: 1) there are no acceptable/unacceptable Description of multiplier components—Martin12 proposed
limits in deformability; and 2) only compression-controlled a simple method to determine multipliers for predicting
failure is considered, although ACI 440.4R-043 states long-term deflections of prestressed concrete structures.
another failure mode (tension-controlled). Overall, a unified That method is outlined below with some modifications to
expression for deformability with specific application limits adjust the original method when necessary. Initial calcula-
is necessary to stipulate the allowable amount of flexural tions for the prestressed concrete deflection multipliers are
deformation in designing FRP-prestressed concrete girders. based on an equation from ACI 318-71,17 describing the
This paper aims to address the aforementioned chal- long-term deflections of reinforced concrete beams, which
lenges in FRP-prestressed concrete technologies: long-term is no longer in use.18 As such, the relationship between long-
multipliers for deflection prediction and deformability. Also term deflections and immediate deflections is adopted from
involved is a study on minimum reinforcement ratio for an updated version. Equation (1) provides the basis for the
FRP-prestressed bridge girders that have been undervalued deflection multiplier (μb) in accordance with ACI 318-1418
in the infrastructure community.
ξ
mb = (1)
RESEARCH SIGNIFICANCE 1 + 50ρ′
ACI Subcommittee 440-I (FRP-Prestressed Concrete)
has recommended several technical items to be updated in
where ξ is a time-dependent constant varying from 1.0 to
a revised version of ACI 440.4R-04.3 Three crucial issues
2.0; and ρ′ is the compression reinforcement ratio at midspan
related to the flexural design of FRP-prestressed members
of the member, regardless of boundary conditions. For the
were identified (long-term multipliers, deformability, and
present study, three reinforcement ratios were used to repre-
minimum reinforcement), because of their largely empir-
sent typical bridge design (that is, ρ′ = 0.001, 0.005, and
ical, ambiguous, and discrepant backgrounds. For instance,
0.01) in tandem with ξ = 2.0 (5 years or more). For long-term
the present long-term multipliers were taken from steel-
deflections from dead load (μdf), the base factor is modified
prestressed concrete despite the fundamental difference
by the strength gain in the concrete, based on the elastic
between FRP and steel materials. Research is conducted to
moduli at release (Eci) and 28 days (Ec), respectively
suggest new design information, which can replace/comple-
ment the existing design articles of ACI 440.4R-04.3
Eci
m df = m b (2)
LONG-TERM MULTIPLIERS Ec
A stochastic simulation is performed to propose long-
term multipliers for AFRP- and CFRP-prestressed concrete Upward displacements for the final camber or deflec-
bridge girders, which can supersede those of ACI 440.4R-043 tions are defined as a function of the prestressing force (μpf).
determined by arbitrary empirical assumptions. Delineated Initial displacements are calculated at release and are modi-
in the following are girder configurations and numerical fied with the final effective prestressing force
approaches that lead to the development of new multipliers.
Pfe
Benchmark bridges m pf = m df (3)
Po
Pursuant to the American Association of State Highway
Transportation Officials (AASHTO) Load and Resistance where Po is the prestressing force immediately following
Factor Design (LRFD) Bridge Design Specifications,15 Kim release; and Pfe is the effective prestressing force. The
and Nickle16 designed 50 prestressed concrete bridges using losses of prestressing force may be determined as detailed in
AFRP and CFRP tendons (25 bridges each). A summary ACI 440.4R-04,3 whose methods are analogous to those of
of the bridge details is provided as follows. Four super- the AASHTO LRFD Bridge Design Specifications.15 Erec-
structure types were employed with variable span lengths tion cambers and deflections may be defined as a portion
and deck widths, as listed in Table 1. These girder config- of the aforementioned final factor multipliers. Martin12
urations were adopted from highway bridges constructed arbitrarily scaled down the μdf factor by 50% to define the
in Colorado, Florida, and Texas. The compressive strength amount of deflection at erection (μde). In the current calibra-
of concrete (fc′) for the girders and decks was from 38 to tion, however, a variable is taken into consideration for the
72 MPa (5.5 to 10.5 ksi) and from 25 to 31 MPa (3.6 to μde factor, which leads to the deflection from dead load at the
4.5 ksi), respectively. The AFRP and CFRP tendons time of erection to be
embedded in the concrete, respectively, had the following
properties: tensile strength = 1250 and 2250 MPa (181 and μde = κμdf (4)
326 ksi); and elastic modulus = 65 and 140 GPa (9400 and
20,600 ksi). The AFRP and CFRP tendons were tensioned to where κ is the erection-adjustment factor to account for creep
and shrinkage at an early stage of the prestressed member.

224 ACI Structural Journal/January 2018


Table 1—Benchmark bridges
ID Type Span, m (ft) Span-depth ratio Girder number Deck width, m (ft) Skew angle, deg
1 B 17 (56) 27 20 25 (82) 0
2 B 17 (56) 27 20 25 (82) 0
3 U 27 (87) 13 4 17 (55) 0
4 B 33 (107) 30 19 27 (90) 40
5 B 33 (108) 30 19 27 (90) 40
6 U 33 (109) 19 4 22 (71) 5
7 U 33 (109) 19 4 22 (71) 5
8 BT 34 (111) 27 7 13 (43) 27
9 BT 36 (119) 23 8 16 (52) 26
10 U 37 (120) 18 4 17 (55) 0
11 U 37 (121) 23 3 9 (29) 8
12 I 38 (124) 21 16 48 (157) 0
13 I 38 (124) 21 16 48 (157) 0
14 B 38 (125) 28 12 20 (67) 0
15 U 38 (125) 19 7 39 (127) 25
16 U 38 (126) 19 8 37 (120) 0
17 B 39 (127) 28 12 20 (67) 0
18 U 39 (127) 19 7 39 (127) 25
19 U 39 (129) 21 8 37 (120) 0
20 B 40 (132) 29 10 13 (43) 56
21 B 41 (135) 30 24 46 (152) 20
22 B 41 (135) 30 24 46 (152) 20
23 U 44 (143) 22 4 17 (55) 0
24 U 44 (144) 22 8 37 (120) 0
25 BT 50 (164) 21 4 9 (28) 30

Notes: ID is identification of bridges; Type (B is box; U is open box; BT is bulb tee; I is I-shape).

Previous research reports that the adjustment factor κ ranges composite multipliers for the load (μdfc) and prestressing
between 0.4 and 0.619; accordingly, random sampling was effects (μpfc), respectively
carried out within this range (details to follow). The camber
component during construction is a multiplier that uses I 
the erection dead load factor as a basis, combined with the m dfc = m de + (m df − m de )  o  (7)
 Ic 
average prestressing force assuming a linear relationship
between pretensioning release and the final effective force.
The upward component from prestressing at erection (μpe) is I 
m pfc = m pe + (m df − m pe )  o  (8)
then defined as  Ic 

Po + Pf e
m pe = m de (5) where Io and Ic are the moments of inertia for noncom-
2 Po posite and composite sections, respectively. For the long-
term deflection of the member induced by placing topping
For the application of long-term superimposed dead loads, concrete, the long-term factor (μt) is expressed as
the base factor may be used to describe the long-term effects
(μsd), as stated in Martin12 I 
m t = m sd  o  (9)
 Ic 
μsd = μb (6)
It is important to note that these factors are intended to be
The use of composite construction is common for precast
additive to the initial deflections calculated; hence, they are
girders. Because composite construction alters the stiffness
increased by 1.0 to be used as direct multipliers, as summa-
of the structural system, the factors shown in Eq. (2) through
rized in Table 2. The factors discussed previously assume
(6) need to be modified. Equations (7) and (8) provide

ACI Structural Journal/January 2018 225


that the composite topping is constructed with a thickness • Statistical distributions: Probability distribution types
of 50 to 75 mm (2 to 3 in.); consequently, the reinforcement are assigned to the variables associated with the sampled
ratio of the system does not change significantly (denoted properties, including statistical characteristics such as
as “with unreinforced composite topping”). By contrast, the coefficients of variation (COVs) shown in Table 3.
most bridge girders are in service with a concrete deck with • Property simulation: Random variables are generated
a noticeable amount of reinforcement, which results in the as per the statistical properties determined in previous
need to consider composite compression reinforcement steps.
ratios (the ρ′ ratios explained previously). For this reason, an • Multiplier sampling based on Monte-Carlo simula-
additional set of multipliers referred to as “with reinforced tion: Using the methodology discussed earlier (Eq. (1)
structural slab” is proposed and determined in which the to (9)) and the variable properties attained, the deflec-
increased reinforcement ratio is accounted. tion multipliers of the noncomposite/composite girder
Random variables and simulation—The multipliers for systems are iteratively simulated until convergence is
long-term deflections enumerated earlier involve several accomplished (that is, the model prediction with random
variables that can be simulated stochastically. A numer- samples approaches a specific value).
ical method called Monte-Carlo simulation that randomly
samples statistical data from a certain probability distri-
bution was used to predict the variability of the bridge
systems. In other words, the simulation technique allows
for the modeling of numerous possible situations in the
performance of individual bridge girders. As an example,
Fig. 1 shows sampled values for concrete strength (fc′) and
prestressing force ratio (Pfe/Po). Following is a summary of
the simulation procedures:
• Property sampling: Technical information is acquired
from the 25 bridges listed in Table 1, including geometric
and material properties that will form the basis for the
statistical means.

Table 2—Deflection multipliers variable summary


with and without composite topping
Stage Component Variable
Deflection (S) 1 + μde
Erection
Camber (P) 1 + μpe
Deflection (S) 1 + μdf
Camber (P) 1 + μpf
Final
Deflection (SD) 1 + μsd
Deflection (CT) 1 + μt*
*
Valid only with composite topping. Fig. 1—Simulated properties for Bridge No. 5: (a) concrete
Notes: Deflection is downward; camber is upward. S is girder self-weight; P is strength fc′; and (b) prestressing force ratio Pfe/Po for AFRP
prestress; SD is superimposed dead load; CT is composite topping. tendons.

Table 3—Random variables


Variable Distribution Mean Bias COV Reference
fci′ Normal Based on material 1.14 to 1.40 0.1 Kwon et al.26
fc′ Normal Based on material 1.14 to 1.40 0.1 Kwon et al.26
Eci Normal 1820√fci′ (in ksi) 1 0.1 AASHTO15
Ec Normal 1820√fci′ (in ksi) 1 0.1 AASHTO15
Efrp Lognormal Based on material — 0.2 Atadero et al.27
Ep Lognormal 28,500 ksi 1.04 0.024 Kwon et al.26
Afrp Normal Based on material — 0.05 Sanchez-Heres et al.28
Ap Normal As designed — 0.0125 Kwon et al.26
d, de, and b Normal As designed — 0.03 Okeil et al.29

Note: 1 ksi = 6.9 MPa.

226 ACI Structural Journal/January 2018


Table 4—Long-term multipliers
AFRP CFRP
Without composite With unreinforced With reinforced Without composite With unreinforced With reinforced
Stage Component topping composite topping structural slab topping composite topping structural slab
Deflection (S) 1.85 1.85 1.85 1.85 1.85 1.85
Erection
Camber (P) 1.70 1.70 1.70 1.80 1.80 1.80
Deflection (S) 2.70 2.35 2.35 2.70 2.35 2.30
Camber (P) 2.00 1.90 1.85 2.55 2.25 2.20
Final
Deflection (SD) 2.90 2.90 2.85 2.90 2.90 2.85
Deflection (CT) — 2.05 2.05 — 2.05 2.05

Notes: Deflection is downward; camber is upward; S is girder self-weight; P is prestress; SD is superimposed dead load; CT is composite topping.

Fig. 3—Comparison of proposed long-term multipliers


against ACI 440.4R-04.
because ACI 440.4R-043 does not specify multipliers with
composite topping. The deflection multipliers at Erection
and Final did not show any difference between the new and
existing multipliers. Substantial dissimilarity was, however,
observed in all other cases, particularly for the multipliers
of camber and deflection (superimposed dead load) at Final.
This discrepancy is attributable to the different material
characteristics between the prestressing steel and the AFRP/
CFRP composite tendons used in the stochastic simulation.
The multipliers suggested in Table 4 verify the qualitative
statement of ACI 440.4R-04,3 which signifies that the multi-
Fig. 2—Calculated multipliers without composite topping: pliers without topping are conservative compared with those
(a) deflection at erection of AFRP-prestressed girders; and with composite topping.
(b) deflection at final of CFRP-prestressed girders.
DEFORMABILITY
• Deflection multipliers: Upon taking the average of The degree of flexural deformation in FRP-prestressed
the simulated multiplier samples, representative deflec- concrete is quantified and assessed to propose a design
tion multipliers are determined for AFRP- and CFRP- expression. The formulation elaborated as follows is valid
prestressed concrete girders with the three topping for bonded applications that satisfy the strain compatibility
conditions: without composite topping, with unrein- of a section. The application of unbonded FRP is not taken
forced composite topping, and with reinforced struc- into account, owing to the wide variability of a strain reduc-
tural slab. tion factor, depending upon the geometry of a prestressed
Recommended multipliers—After solving multipliers member.
for all 50 bridge cases with AFRP and CFRP tendons, the
results were sorted, as shown in Fig. 2. For design purposes, Derivation
conservative multipliers were selected and proposed (Table Equation (10) defines a newly proposed deformability
4). Figure 3 compares the proposed multipliers with the index (DI) for an FRP-prestressed member
existing ones in ACI 440.4R-043 for both AFRP- and CFRP-
prestressed bridge girders. It should be noted that the compar- ε ultf
ison focuses only on the cases without composite topping, DI = (10)
ε crf

ACI Structural Journal/January 2018 227


where εfult and εfcr are the FRP strains at ultimate and Table 5—Lower limits of deformability index (DI)
cracking of the member, respectively. In accordance with Condition AFRP CFRP
strain compatibility for a bonded section, these prestressed
Tension-controlled 2.8 1.8
FRP strains consist of three components
Compression-controlled 1.5 1.5
εfult = εfe + εce + εfail (11)
ε fe + (d f /c − 1)ε cu
εfcr = εfe + εce + εfcr (12) DI = for compression-controlled section
(d f − c) f r
ε fe +
εfe = ffe/Ef (13) h − c Ef
(17b)
Pfe ( I g + e 2f Ag )
ε ce = (14) Equations (17a) and (17b) are further manipulated for
Ag I g Ec
practical implementation. Because the modulus of rupture
of concrete is much smaller than the tensile modulus of FRP,
where εfe is the effective FRP strain after prestress losses; the fr/Ef term is ignored. Substituting (df/c –1) for k gives
εce is the additional FRP strain when the cambered member
is decompressed (concrete strain at the FRP level becomes ε fu
zero under flexural loading, which results in the neutral posi- DI = for tension-controlled section (18a)
tion of the cross section); εfail and εfcr are the FRP strains ε fe
measured from the neutral position to failure and to cracking,
respectively; ffe and Ef are the effective stress and modulus of k ε cu
the FRP, respectively; Ig and Ag are the moment of inertia and DI = for compression-controlled section (18b)
ε fe
cross-sectional area of the uncracked member, respectively;
and ef is the distance from the neutral axis of the section to
the center of gravity of the FRP. The εfail component is deter- Application
mined as per the failure mode of the prestressed member, Conventional prestressed concrete beams can sufficiently
including the conventional similar triangles of strain devel- deform under flexural loading, if the following limit is
opment in compression and tension achieved20

εfail = εfu for tension-controlled section (15a) c


≤ 120ε cu (19)
h
(d f − c)
ε fail = ε cu for compression-controlled section (15b) Given that prestressed concrete members should exhibit
c acceptable flexural deformation, regardless of prestressing
material, Eq. (19) can be used for FRP-prestressed concrete.
where εfu is the ultimate FRP strain; c and df are the neutral Substituting Eq. (19) into Eq. (18b) yields
axis depth of the section and the effective depth of the FRP,
respectively, measured from the extreme compression fiber d f / (120h) − ε cu + ε fe
of the member; and εcu is the maximum useable concrete DI ≥ (20)
ε fe
strain (εcu = 0.003 [from Reference 18]). The εfcr component
is also attained from the similar triangles of the uncracked For a typical prestressed concrete section, the ratio
section’s stress profile between the effective depth and height of the section is
0.85.20 ACI 440.4R-043 specifies that AFRP and CFRP
(d f − c) f r tendons are tensioned up to 0.4ffu and 0.6ffu, respectively,
ε crf = (16)
h − c Ef including transfer losses. Because the long-term prestress
losses of FRP-prestressed concrete members may be esti-
where h is the height of the member; and fr is the modulus of mated by the methods for steel-prestressed concrete,3 the
rupture of concrete (fr = 7.5√fc′ in psi and 0.63√fc′ in MPa18). lump-sum prestress losses provided in the AASHTO LRFD
Combining Eq. (10) to (16) with the fact that the εce term is Bridge Design Specifications15 are adopted, which are less
negligible in most cases, the following expressions are obtained than 10% of the ultimate tendon strength. As a result, the
long-term prestress losses of FRP-prestressed concrete are
ε fu assumed to be 0.05ffu. It is worth noting that this long-term
DI = for tension-controlled section (17a) loss assumption is reasonable because FRP-prestressed
(d f − c) f r
ε fe + concrete experiences less prestress losses than its steel coun-
h − c Ef
terpart, due to low FRP modulus.3 Considering all these
components, the effective strains of AFRP and CFRP may

228 ACI Structural Journal/January 2018


the deformability of tension-controlled sections was not
evaluated with these girders, because the index can readily
be attained using Eq. 18(a) without rigorous computation.
A wide range of prestressing levels (immediately after
transfer) were considered, varying from 0.1ffu up to 0.4ffu
and 0.6ffu for the AFRP and CFRP tendons, respectively, in
compliance with the prestressing limits of ACI 440.4R-04,3
as shown in Fig. 4(b).
Flexural behavior—Conventional sectional analysis
was conducted to predict the moment-curvature behavior
of the benchmark girders at midspan. Typical trilinear and
bilinear responses are provided in Fig. 5(a) for the steel- and
AFRP-prestressed BT-72 girders, respectively. Because of
low prestressing levels, the cracking moments of the AFRP
girders were lower than that of the steel girder (for example,
the cracking moments of BT-72 prestressed with AFRP at
0.4ffu and 0.1ffu were 26% and 72% lower, respectively).
With a reduction in prestressing level of the AFRP tendon,
the girders’ energy dissipatiion up to failure increased
(Fig. 5(a), inset), which was obtained by integrating the
area under the individual moment-curvature curve. Figure
5(b) exhibits the variation of a ratio between the neutral axis
depth and height of the girders at failure (c and h, respec-
tively). While the c/h ratios went up with a prestressing
level, irrespective of FRP and structural types, the girders
with CFRP revealed higher ratios than those with AFRP.
The reason is that the CFRP’s stress development was not as
susceptible to girder bending as was the AFRP’s; hence, the
neutral axes of the CFRP girders were consistently deeper.
The c/h ratios of all girders were within the limit of 120εcu
(Eq. (19)). This indicates that the girders had sufficiently
deformed until failure occurred, even though the curva-
tures of the AFRP-prestressed girders were less developed
Fig. 4—Prestressed concrete girder details: (a) dimension in
(lesser curvature) in comparison with the curvature of the
mm (midspan); and (b) prestressing properties. (Note: 1 mm =
steel-prestressed girder (Fig. 5(a)). The ratios between the
0.0394 in.; 1 kN = 0.225 kip.)
ultimate and cracking moments of the girders with AFRP/
be 35% and 55% of their ultimate strains, respectively. CFRP (Mn and Mcr, respectively) are plotted in Fig. 5(c),
Following ACI 440.4R-04,3 the lower bound of ultimate including those of BT-72 and BIII-48 with steel for compar-
strains in typical AFRP and CFRP tendons may be taken as ison (the constant Mn/Mcr ratios of steel are attributable to
0.015. The limits of the deformability indices provided in the behavior of the steel-prestressed case; Fig. 5(a)). As a
Eq. (18) are then calculated, and summarized in Table 5 with prestressing level in the AFRP/CFRP tendons increased, the
rounded values for design convenience. Mn/Mcr ratios asymptotically decayed down to the ratios of
the steel-prestressed cases. The BT-72 girders maintained
Assessment higher Mn/Mcr ratios compared with the BIII-48 girders in
Benchmark girders—Two common prestressed concrete all tendon types due to the former’s longer moment arm
bridge girders were selected from the PCI Bridge Design (BIII-48’s height is 53% of BT-72’s; Table 6), which caused
Manual21 to assess the deformability index proposed earlier: higher moment-carrying capacity Mn. Figure 5(d) shows
AASHTO Box Beam BIII-48 and AASHTO/PCI Bulb Tee deformability indexes for the AFRP- and CFRP-prestressed
BT-72. Figure 4(a) and Table 6 show the girders’ geometric girders. The decreasing trend of deformability was similar to
and material properties. For numerical parametric investiga- that of the Mn/Mcr ratios, whereas there was no remarkable
tions, the BIII-48 and BT-72 girders were used as the default difference between the BT-72 and BIII-48 girders. It is there-
sections with representative AFRP and CFRP tendons fore stated that the proposed deformability index (Eq. (18a)
whose properties were taken from ACI 440.4R-04,3 as listed and (18b)) can be universally applicable to bridge girders
in Table 6. The amounts of AFRP and CFRP tendons were prestressed with either AFRP or CFRP tendons.
determined to provide the flexural capacities of the sections Parametric study—To evaluate the effects of a potential
as close as possible to the capacities of the steel-prestressed change in FRP modulus on the deformability of prestressed
sections, with a condition that tendon stresses did not exceed concrete girders, a parametric study was conducted using
their ultimate strengths when the girders failed by concrete BT-72 and BIII-48 girders with the preceding AFRP and
crushing (compression-controlled). It should be noted that CFRP tendons. This investigation is meaningful from appli-

ACI Structural Journal/January 2018 229


Table 6—Geometric and material properties of prestressed concrete members
Property Box Beam (BIII-48) Bulb Tee (BT-72)
Height (h) 990 mm (39 in.) 1829 mm (72 in.)
Effective depth (d) 874 mm (34.42 in.) 1653 mm (65.08 in.)
Cross-sectional area of section (Ac) 524,515 mm (813 in. )
2 2
494,838 mm2 (767 in.2)
Moment of inertia (I) 70.2 × 109 mm4 (168,367 in.4) 227 × 109 mm4 (545,894 in.4)
Tendon eccentricity at midspan (e) 374 mm (14.71 in.) 754 mm (29.68 in)
Distance from centroid to top of section (yt) 500 mm (19.71 in.) 899 mm (35.4 in.)
Distance from centroid to bottom of section (yb) 490 mm (19.29 in.) 930 mm (36.6 in.)
Radius of gyration (r) 366 mm (14.4 in.) 678 mm (26.7 in.)
Concrete strength at initial prestress (fci′) 28 MPa (4000 psi) 40 MPa (5800 psi)
Specified concrete strength at 28 days (fc′) 34 MPa (5000 psi) 45 MPa (6500 psi)
Concrete elastic modulus at transfer (Eci) 25 GPa (3600 ksi) 30 GPa (4340 ksi)
Concrete elastic modulus (Ec) 28 GPa (4030 ksi) 32 GPa (4595 ksi)
Area of prestressing steel (Ap) 2863 mm2 (4.437 in.2) 4738 mm2 (7.344 in.2)
Area of AFRP tendon (AAFRP) 4516 mm2 (7 in.2) 8387 mm2 (13 in.2)
Area of CFRP tendon (ACFRP) 2581 mm2 (4 in.2) 4194 mm2 (6.5 in.2)
Elastic modulus of prestressing steel (Ep) 197 GPa (28,500 ksi) 197 GPa (28,500 ksi)
Elastic modulus of AFRP tendon (EAFRP) 62 GPa (8992 ksi) 62 GPa (8992 ksi)
Elastic modulus of CFRP tendon (ECFRP) 150 GPa (21,756 ksi) 150 GPa (21,756 ksi)
Ultimate strength of prestressing steel (fpu) 1860 MPa (270 ksi) 1860 MPa (270 ksi)
Yield strength of prestressing steel (fpy) 1675 MPa (243 ksi) 1675 MPa (243 ksi)
Ultimate strength of AFRP tendon (fAFRPu) 1200 MPa (174 ksi) 1200 MPa (174 ksi)
Ultimate strength of CFRP tendon (fCFRPu) 2550 MPa (370 ksi) 2550 MPa (370 ksi)

Fig. 5—Flexural behavior of prestressed concrete members: (a) moment curvature of BT-72 with AFRP; (b) ratio between
neutral axis depth and girder height at failure of girders; (c) ratio between ultimate and cracking moments; and (d) assessment
of deformability. (Note: 1 kN·m = 738 lb-ft.)

230 ACI Structural Journal/January 2018


Fig. 6—Effect of FRP modulus on deformability: (a) moment curvature of bulb tee (BT-72) with CFRP tendons; (b) variation
of FRP strain; (c) deformability index of bulb tee (BT-72); and (d) deformability index of box beam (BIII-48).
cation standpoints because FRP-prestressed concrete girders
are designed without considering any possibility of altered
FRP properties3 that may take place within a service period
of 75 years. Although FRP composites are known as durable
materials, recent research shows that their elastic moduli
vary with time when subjected to environmental/mechanical
distress.22,23 According to experimental observations,22,23
the variation range of FRP modulus was determined to be
±10% of the nominal value. Additionally, it was assumed
that the rupture strain of FRP is invariant with time as per the
time-dependent strain failure criteria of polymeric composite
materials.24
Figure 6(a) exhibits the flexural responses of CFRP- Fig. 7—Assessment of deformability using benchmark bridges.
prestressed BT-72 in conjunction with variable moduli from that the deformability of FRP-prestressed girders would
90%Efo to 110%Efo, where Efo is the nominal modulus of not cause structural problems, as long as: 1) the girders
the CFRP (Table 6). The precracking behavior of the girder are properly designed per the proposed design recommen-
was not influenced by modulus variations; however, the dations; and 2) the tendons’ actual moduli are greater than
cracking moment was affected. Because the contribution those used in design (this is generally not a concern, because
of the CFRP modulus to the concrete girder’s stiffness was manufacturers’ recommended properties are conservative).
negligible, the post-cracking stiffness of the girder remained It is, however, important to note that care should be exer-
unchanged. An interesting finding is that the reduced CFRP cised when FRP’s modulus is expected to degrade in certain
modulus caused a transition in the girder’s failure mode from service environments that can alter the failure mode of the
compression-controlled to tension-controlled, as shown in prestressed girders, as mentioned earlier.
Fig. 6(b). This is attributed to the fact that the decreased
modulus brought about an increase in CFRP strain under Applicability
the same level of tensile stress. Figures 6(c) and 6(d) reveal The deformability indexes of the full-scale prestressed
the effects of CFRP modulus on deformability of the BT-72 concrete bridges with AFRP and CFRP tendons (Table 1)
and BIII-48 girders with AFRP and CFRP tendons, respec- were calculated to examine the applicability of the proposed
tively. For both girders, the deformability indexes gradu- limits (Table 5), as shown in Fig. 7. Because the low-mod-
ally dwindled with the increased modulus of the tendons. ulus AFRP was more deformable in principle, the indexes of
Nonetheless, the indexes of all compression-controlled cases the bridges with AFRP were higher than those of the bridges
were higher than the limit of 1.5. These observations denote with CFRP. Unlike the bridges with compression-controlled

ACI Structural Journal/January 2018 231


Table 7—Ratio between ultimate and cracking capacities
Experimental* Experimental*
Reference Cracking Ultimate Failure †
Reference Cracking Ultimate Failure†
G&S 643.9 kN 2,443 kN T N&T 11,454 Nm 27,199 Nm T
Stoll 178 kN 554 kN T N&T 12,850 Nm 26,905 Nm T
T&S 85 kN 195 kN T N&T 12,256 Nm 26,932 Nm C
T&S 85 kN 185 kN T N&T 11,825 Nm 27,687 Nm C
T&S 86 kN 155 kN T N&T 12,531 Nm 25,450 Nm C
T&S 84.7 kN 143 kN T N&T 8821 Nm 15,398 Nm T
N&T 5852 Nm 12,311 Nm T N&T 7729 Nm 14,635 Nm T
N&T 6239 Nm 12,243 Nm T N&T 8552 Nm 20,020 Nm C
N&T 5152 Nm 11,955 Nm T N&T 8402 Nm 20,625 Nm C
N&T 5818 Nm 14,607 Nm T N&T 6054 Nm 19,114 Nm T
N&T 5805 Nm 15,381 Nm T N&T 13,265 Nm 18,710 Nm T
N&T 6720 Nm 17,197 Nm C N&T 12,902 Nm 21,069 Nm T
N&T 6080 Nm 18,681 Nm C N&T 12,006 Nm 20,541 Nm T
N&T 5555 Nm 13,442 Nm T M&E 17.2 kNm 34.8 kNm T
N&T 5086 Nm 12,424 Nm T M&E 17.2 kNm 33.2 kNm T
N&T 6112 Nm 12,516 Nm T M&E 18.1 kNm 40 kNm T
N&T 6112 Nm 11,948 Nm T Niitani 55 kN 131 kN C
N&T 5490 Nm 14,561 Nm C Niitani 54 kN 129 kN C
N&T 5792 Nm 14,386 Nm T Niitani 54 kN 103 kN C
N&T 5871 Nm 14,557 Nm T T&M 7 kN 19 kN C
N&T 5331 Nm 15,257 Nm T T&M 9 kN 17.9 kN C
N&T 6759 Nm 16,295 Nm T T&M 9 kN 20.9 kN T
N&T 6613 Nm 16,911 Nm T T&M 8 kN 15 kN C
N&T 8821 Nm 23,388 Nm T T&M 8 kN 22.5 kN C
N&T 9267 Nm 24,306 Nm T T&M 8 kN 28 kN C
N&T 12,087 Nm 27,317 Nm T T&M 9 kN 18 kN C
*
Flexural moment or load.

T is tension-controlled; C is compression-controlled.
Notes: G&S is Grace and Singh30; Stoll is Stoll et al.31; T&S is Toutanji and Saafi32; N&T is Nanni and Tanigaki5; M&E is McKay and Erki33; Niitani is Niitani et al.34; T&M is
Taerwe and Mattys35; 1 kN = 0.225 kip; 1 Nm = 0.74 ft-lb; 1 kNm = 740 ft-lb.

sections satisfying the deformability limit, several cases with ratios (Fig. 8(a)). The average ratios of the beams failed by
tension-controlled sections demonstrated the indexes lower concrete crushing and FRP rupture were 2.3 and 2.4, respec-
than corresponding limits (Table 5), including maximum tively. Figure 8(b) shows the factored ratios of the experi-
margins of 4% and 5% for the AFRP- and CFRP-prestressed mental responses. The strength reduction factors were taken
bridges, respectively. This issue is attributable to the fact from Kim and Nickle,16 which were calibrated using reli-
that the bridges were initially designed without considering ability theory, rather than the empirically determined reduc-
deformability16; nonetheless, these insignificant margins tion factors of ACI 440.4R-04.3 In all cases except one beam
can still support that the proposed limits are applicable in failed by bond slip, the factored ratios were greater than
practice (the margins for the tension-controlled sections can 1.2, including average ratios of 1.7 and 1.9 for the tension-
readily be addressed by slightly adjusting the amounts of the controlled and compression-controlled beams, respectively.
AFRP and CFRP tendons). While these ratios were obtained from laboratory-scale
beams, their variation range was in reasonable agree-
MINIMUM REINFORCEMENT ment with the range of the full-scale girders (Fig. 5(c) and
Table 7 summarizes experimental data collated from Fig. 8(c) and 8(d)). The ratio between a factored flexural
published literature on FRP-prestressed concrete beams. resistance (ϕMn) and its cracking counterpart (Mcr) is there-
To account for the significance of failure modes, either fore proposed to be greater than 1.2, which is identical to
compression- or tension-controlled, the experimental data the ratio specified in ACI 318-14,18 the PCI Design Hand-
were sorted and presented in terms of ultimate-to-cracking book,25 and AASHTO LRFD Bridge Design Specifica-

232 ACI Structural Journal/January 2018


Fig. 8—Ratio between ultimate and cracking responses: (a) unfactored (experimental); (b) factored (experimental): (c) unfac-
tored (benchmark full-scale bridges); and (d) factored (benchmark full-scale bridges).
tions.15 It should be noted that ACI 440.4R-043 states a compression-controlled sections. The minimum reinforce-
ϕMn/Mcr ratio of 1.5 that was arbitrarily elevated from the ment requirement on the cracking stage of prestressed
ratio of 1.2 used in steel-prestressed concrete sections to concrete girders was revisited for both improvement and
avoid a concern associated with a tension-controlled FRP consistency with other design specifications. The following
section, which does not appear to be an issue considering the conclusions are drawn:
high strength of FRP composites (in other words, adequately 1. The proposed long-term multipliers for AFRP/
designed FRP-prestressed concrete girders will not fail CFRP-prestressed concrete girders were the same as those of
abruptly immediately after cracking, regardless of failure ACI 440.4R-043 for deflections at erection and final stages,
mode that is related to the ultimate limit state rather than at whereas the multipliers were different in all other cases
a cracking stage). Further, because the amount of minimum because of the AFRP/CFRP’s unique material characteristics
reinforcement is not necessarily engaged with FRP types, associated with the stochastic calibration.
the proposed ϕMn/Mcr ratio of 1.2 for AFRP- and CFRP- 2. Although the moment-curvature behavior of the AFRP/
prestressed members can be integrated in design. CFRP-prestressed girders was not comparable with that
of steel-prestressed counterparts, the girders sufficiently
SUMMARY AND CONCLUSIONS deformed prior to failure in either concrete-crushing or
This paper has addressed technical challenges regarding FRP-rupture. With an increase in the prestressing level, the
FRP-prestressed concrete bridge girders, particularly for deformability indexes of the girders with AFRP and CFRP
those of ACI 440.4R-043: long-term multipliers, deform- were asymptotically reduced toward the suggested limits.
ability index, and minimum reinforcement. New design 3. The alteration of the AFRP/CFRP moduli influenced the
information was proposed based on analytical derivations girders’ cracking moments; however, corresponding effects
along with an experimental assessment. Fifty previously- on precracking behavior and post-cracking stiffness were
designed highway bridges prestressed with AFRP and CFRP negligible. The degraded moduli, however, changed the
tendons were also employed to evaluate the proposed design failure mode of the girders from compression-controlled to
expressions. The long-term multipliers were refined into tension-controlled, which is an important consideration in
three categories for AFRP and CFRP applications at Erec- aggressive service environments.
tion and Final without composite topping, with unreinforced 4. The average ultimate-to-cracking ratios of the 52 exper-
composite topping, and with reinforced structural slab. These imental AFRP/CFRP-prestressed concrete beams were 2.3
are different from the single category (without composite and 2.4 for the tension-controlled and compression-con-
topping) specified in ACI 440.4R-043 that was taken from trolled sections, respectively, which were in agreement with
the empirically-determined values with steel strands. The the ratios of the full-scale bridge girders. The factored ulti-
new deformability index with specific limits was appraised mate-to-cracking moment ratio of 1.2 proposed was suffi-
in two possible failure conditions: tension-controlled and cient to provide the low bound for AFRP/CFRP-prestressed

ACI Structural Journal/January 2018 233


concrete, and is in conformance with the design expres- 14. Au, F. T. K., and Du, J. S., “Deformability of Concrete Beams with
Unbonded FRP Tendons,” Engineering Structures, V. 30, No. 12, 2008,
sion of ACI 318-14,18 the PCI Design Handbook,25 and the pp. 3764-3770. doi: 10.1016/j.engstruct.2008.07.003
AASHTO LRFD Bridge Design Specifications.15 15. AASHTO, “AASHTO LRFD Bridge Design Specifications,” seventh
edition with interim revisions, American Association of State Highway and
Transportation Officials, Washington, DC, 2016.
AUTHOR BIOS 16. Kim, Y. J., and Nickle, R. W., “Strength Reduction Factors for Fiber-
Yail J. Kim, FACI, is a Professor in the Department of Civil Engi-
Reinforced Polymer Prestressed Concrete Bridges in Flexure,” ACI
neering at the University of Colorado Denver, Denver, CO. He is Chair
Structural Journal, V. 113, No. 5, Sept.-Oct. 2016, pp. 1043-1052. doi:
of ACI Committee 345, Concrete Bridge Construction, Maintenance, and
10.14359/51689028
Repair; and ACI Subcommittee 440-I, FRP-Prestressed Concrete. He is
17. ACI Committee 318, “Building Code Requirements for Reinforced
a member of ACI Committees 342, Evaluation of Concrete Bridges and
Concrete (ACI 318-71),” American Concrete Institute, Farmington Hills,
Bridge Elements; 343, Concrete Bridge Design; and 440, Fiber-Reinforced
MI, 1971, 78 pp.
Polymer Reinforcement. His research interests include advanced composite
18. ACI Committee 318, “Building Code Requirements for Structural
materials for structural rehabilitation, complex systems, uncertainty quan-
Concrete (ACI 318-14) and Commentary (ACI 318R-14),” American
tification, and science-based structural engineering, including statistical,
Concrete Institute, Farmington Hills, MI, 2014, 520 pp.
interfacial, and quantum physics.
19. ACI Committee 209, “Prediction of Creep, Shrinkage, and Tempera-
ture Effects in Concrete Structures,” Designing for Effects of Creep,
Raymon W. Nickle is a Structural Engineer at AECOM. He received his BS
Shrinkage and Temperature in Concrete Structures, SP-27, American
and MS in civil engineering from Colorado State University, Fort Collins,
Concrete Institute, Farmington Hills, MI, 1971, pp. 51-93.
CO, in 2009 and the University of Colorado Denver in 2015, respectively.
20. Skogman, B. C.; Tadros, M. K.; and Grasmick, R., “Ductility of
His research interests include the design and analysis of transportation
Reinforced and Prestressed Concrete Flexural Members,” PCI Journal,
structures.
V. 33, No. 6, 1988, pp. 94-107. doi: 10.15554/pcij.11011988.94.107
21. PCI, PCI Bridge Design Manual, Precast/Prestressed Concrete Insti-
ACKNOWLEDGMENTS tute, Chicago, IL, 2003.
The authors gratefully acknowledge the members of ACI 440-I Subcom- 22. Karbhari, V. M., “Durability Data for FRP Rehabilitation Systems,”
mittee, FRP-Prestressed Concrete, chaired by the first author of this paper, Report No. CA09-0248, California Department of Transportation, Sacra-
for their rigorous review of technical contents and valuable suggestions. mento, CA, 2009.
Proprietary information such as product names and manufacturers is not 23. Lees, J. M.; Toumpanaki, E.; Barbezat, M.; and Terrasi, G. P.,
included herein. “Mechanical and Durability Screening Test Methods for Cylindrical CFRP
Prestressing Tendons,” Journal of Composites for Construction, ASCE, V.
21, No. 2, 2017, p. 04016080 doi: 10.1061/(ASCE)CC.1943-5614.0000727
REFERENCES 24. Guedes, R. M., “Time-Dependent Failure Criteria for Polymer Matrix
1. Abdelrahman, A. A.; Tadros, G.; and Rizkalla, S. H., “Test Model for Composites: A Review,” Journal of Reinforced Plastics and Composites,
the First Canadian Smart Highway Bridge,” ACI Structural Journal, V. 92, V. 29, No. 20, 2010, pp. 3041-3047. doi: 10.1177/0731684410370067
No. 4, July-Aug. 1995, pp. 451-458. 25. PCI, PCI Design Handbook, seventh edition, Precast/Prestressed
2. Fam, A. Z.; Rizkalla, S. H.; and Tadros, G., “Behavior of CFRP for Concrete Institute, Chicago, IL, 2010.
Prestressing and Shear Reinforcements of Concrete Highway Bridges,” 26. Kwon, O.-S.; Kim, E.; Orton, S.; Salim, H.; and Hazlett, T., “Cali-
ACI Structural Journal, V. 94, No. 1, Jan.-Feb. 1997, pp. 77-86. bration of the Live Load Factor in LRFD Design Guidelines,” Final Report
3. ACI Committee 440, “Prestressing Concrete Structures with FRP No. OR11-003, Missouri Department of Transportation, Jefferson City, MO,
Tendons (ACI 440.4R-04),” American Concrete Institute, Farmington 2011.
Hills, MI, 2004, 35 pp. 27. Atadero, R. A., and Karbhari, V. M., “Calibration of Resistance
4. Youakim, S. A., and Karbhari, V. M., “An Approach to Determine Factors for Reliability-Based Design of Externally-Bonded FRP Compos-
Long-Term Behavior of Concrete Members Prestressed with FRP Tendons,” ites,” Composites. Part B, Engineering, V. 39, No. 4, 2008, pp. 665-679.
Construction and Building Materials, V. 21, No. 5, 2007, pp. 1052-1060. doi: 10.1016/j.compositesb.2007.06.004
doi: 10.1016/j.conbuildmat.2006.02.006 28. Sanchez-Heres, L. F.; Ringsberg, J. W.; and Johnson, E., “Influence
5. Nanni, A., and Tanigaki, M., “Pretensionded Prestressed Concrete of Mechanical and Probabilistic Models on the Reliability Estimates of
Members with Bonded Fiber Reinforced Plastic Tendons: Development Fibre-Reinforced Cross-Ply Laminates,” Structural Safety, V. 51, 2014,
and Flexural Bond Lengths (Static),” ACI Structural Journal, V. 89, No. 4, pp. 35-46. doi: 10.1016/j.strusafe.2014.06.001
July-Aug. 1992, pp. 433-441. 29. Okeil, A.; El-Tawil, S.; and Shahawy, M., “Flexural Reliability of
6. Grace, N. F., and Abdel-Sayed, G., “Behavior of Carbon Fiber- Reinforced Concrete Bridge Girders Strengthened with Carbon Fiber-
Reinforced Prestressed Concrete Skew Bridges,” ACI Structural Journal, Reinforced Polymer Laminates,” Journal of Bridge Engineering, ASCE,
V. 97, No. 1, Jan.-Feb. 2000, pp. 26-35. V. 7, No. 5, 2002, pp. 290-299. doi: 10.1061/(ASCE)1084-0702(2002)7:5(290)
7. Dolan, C. W., and Swanson, D., “Development of Flexural Capacity 30. Grace, N. F., and Singh, S. B., “Design Approach for Carbon
of an FRP Prestressed Beam with Vertically Distributed Tendons,” Fiber-Reinforced Polymer Prestressed Concrete Bridge Beams,” ACI Struc-
Composites. Part B, Engineering, V. 33, No. 1, 2002, pp. 1-6. doi: 10.1016/ tural Journal, V. 100, No. 3, May-June 2003, pp. 365-376.
S1359-8368(01)00053-1 31. Stoll, F.; Saliba, J. E.; and Casper, L. E., “Experimental Study of
8. Saiedi, R.; Fam, A.; and Green, M. F., “Behavior of CFRP-Prestressed CFRP-Prestressed High-Strength Concrete Bridge Beams,” Composite
Concrete Beams under High-Cycle Fatigue at Low Temperature,” Journal Structures, V. 49, 2000, pp. 191-200. doi: 10.1016/S0263-8223(99)00134-8
of Composites for Construction, ASCE, V. 15, No. 4, 2011, pp. 482-489. 32. Toutanji, H., and Saafi, M., “Performance of Concrete Beams
doi: 10.1061/(ASCE)CC.1943-5614.0000190 Prestressed with Aramid Fiber-Reinforced Polymer Tendons,”
9. Zhang, K.; Fang, Z.; and Nanni, A., “Behavior of Tendons with Composite Structures, V. 44, No. 1, 1999, pp. 63-70. doi: 10.1016/
Multiple CFRP Rods,” Journal of Structural Engineering, ASCE, V. 142, S0263-8223(98)00126-3
No. 10, 2016, p. 04016065. doi: 10.1061/(ASCE)ST.1943-541X.0001535 33. McKay, K. S., and Erki, M. A., “Flexural Behaviour of Concrete
10. Soudki, K. A., “FRP Reinforcement for Prestressed Concrete Struc- Beams Pretensioned with Aramid Fibre Reinforced Plastic Tendons,”
tures,” Progress in Structural Engineering and Materials, V. 1, No. 2, 1998, Canadian Journal of Civil Engineering, V. 20, No. 4, 1993, pp. 688-695.
pp. 135-142. doi: 10.1002/pse.2260010205 doi: 10.1139/l93-085
11. Lees, J. M., “Fibre-Reinforced Polymers in Reinforced and Prestressed 34. Niitani, K.; Tezuka, M.; and Tamura, T., “Flexural Behavior of
Concrete Applications: Moving Forward,” Progress in Structural Engi- Prestressed Concrete Beams Using AFRP Pre-Tensioning Tendons,”
neering and Materials, V. 3, No. 2, 2001, pp. 122-131. doi: 10.1002/pse.60 Proceedings of the 3rd International Symposium on Non-Metallic (FRP)
12. Martin, L. D., “A Rational Method for Estimating Camber and Reinforcement for Concrete Structures, 1997, pp. 663-670.
Deflection of Precast Prestressed Members,” PCI Journal, V. 22, No. 1, 35. Taerwe, L., and Matthys, S., “Structural Behavior of Concrete Slabs
1977, pp. 100-108. Prestressed with AFRP Bars,” Proceedings of the 2nd International Sympo-
13. Zou, P. X. W., “Flexural Behavior and Deformability of Fiber Rein- sium on Non-Metallic (FRP) Reinforcement for Concrete Structures, 1995,
forced Polymer Prestressed Concrete Beams,” Journal of Composites pp. 421-429.
for Construction, ASCE, V. 7, No. 4, 2003, pp. 275-284. doi: 10.1061/
(ASCE)1090-0268(2003)7:4(275)

234 ACI Structural Journal/January 2018


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 115-S19

Seismic Performance of Reinforced Concrete Columns


with Lap Splices in Plastic Hinge Region
by Chul-Goo Kim, Hong-Gun Park, and Tae-Sung Eom
For convenient bar placement, longitudinal bars in columns are
often lap-spliced at the bottom of the columns where plastic hinge
forms. In this study, the seismic performance of lap-spliced columns
was investigated through cyclic loading tests. The bar splice length
and shear span-to-height ratio were considered as the primary
test variables. The test results showed that after flexural yielding,
bond-splitting cracks propagated along the lap-spliced bars during
repeated load cycles. The strength, ductility, and failure mode of
the columns were significantly affected by the bar splice length
and shear span-to-height ratio (or moment gradient). Based on the
results, effects of moment gradient on the bond demand of spliced
bars were studied, and a reduced splice length addressing the
moment gradient was proposed. Furthermore, recommendations
for the design and detailing of lap splices in the column plastic
hinge region such as splice length, bar offset, and confining trans-
verse reinforcement were provided.

Keywords: cyclic test; lap splice; offset bar; reinforced concrete column;
seismic performance.
Fig. 1—Location of lap splices in reinforced concrete
columns.
INTRODUCTION
The seismic performance of reinforced concrete (RC) [58 ksi] and ls/db = 47.8 ~ 79.6 (ACI 318-14) and 35.1 ~ 71.5
columns such as ductility and energy dissipation capacity (Eurocode 2) for fy = 500 MPa (72.5 ksi).
can be degraded when lap splices of longitudinal reinforcing In previous studies on lap splices of column longitu-
bars are located in the plastic hinge zone. For this reason, dinal bars, Lynn et al.3 investigated the cyclic behavior of
in ACI 318-14,1 lap splices are not permitted in potential columns with lap-spliced bars (ls = 20db). The test results
plastic hinge zones of special moment frames. On the other showed that the column behavior was significantly affected
hand, such a requirement is not imposed for ordinary and by the spacing s of transverse hoops in the lap-splice region:
intermediate moment frames. Thus, in low and moderate the column behavior was ductile for s = 0.67hmin (305 mm
seismic regions, for convenience of bar placement, lap [12 in.]), but brittle for s = hmin (457 mm [18 in.]). Aboutaha
splices of longitudinal bars are usually located at the bottom et al.4 reported that in columns with lap-spliced bars of ls =
of columns, which is a plastic hinge zone (Fig. 1). 24db, bond-splitting failure occurred along the length of
When lap splices of longitudinal bars are used at the plastic the lap splice before flexural yielding. Melek and Wallace5
hinge zone, the strength and ductility of columns are affected performed cyclic loading tests of columns with ls = 20db and
by the lap-splice length. Figure 2 shows the required tension s = 0.67hmin (305 mm [12 in.]). Brittle failure occurred in all
splice length ls specified in ACI 318-141 and Eurocode 2.2 The columns at relatively small lateral drift ratios of 1.0% ~ 1.5%
horizontal and vertical axes denote the concrete compressive because of the short lap-splice length. Similarly, Haroun and
strength (fc′ = 18 ~ 30 MPa [2.61 ~ 4.35 ksi]) and required Elsanadedy,6 Harries et al.,7 and Harajli8 reported that brittle
splice length-to-bar diameter ratio (ls/db), respectively. In this bond failure occurred at lateral drift ratios of 1.0% ~ 1.5% in
comparison, the lap-splice length specified in ACI 318-14 is columns with ls = 20db, 22db, and 30db.
defined as ls = 1.3ld = 1.3(0.9dbfy/√fc′)/([cb + Ktr]/db) for Splice Furthermore, the strength and ductility of columns are
Class B. In Eurocode 2, the lap-splice length is defined as affected by lap-splice details. ACI 3159 specifies two
ls = α6lbd = α2α3α6(dbfyd)/(4fbd) by assuming α1 = α4 = α5 = 1.0, offset-bar details, bottom and top offset-bar splices (refer
where fyd is the design yield stress of the spliced bar, fbd is the to Fig. 3). In the bottom offset-bar splice, the bottom bars
concrete bond strength (= 2.25fctd = 0.47fck2/3), and α2 (= 0.7 ~ from the lower story are offset inside. In the top offset-bar
1.0) and α3 (= 0.7 ~ 1.0) are the coefficients for concrete cover
and confining transverse reinforcement, respectively. If the ACI Structural Journal, V. 115, No. 1, January 2018.
lap splices of column longitudinal bars are placed at the same MS No. S-2017-071, doi: 10.14359/51701109, was received February 27, 2017, and
reviewed under Institute publication policies. Copyright © 2018, American Concrete
location (α6 = 1.5), the splice lengths are ls/db = 38.2 ~ 63.7 Institute. All rights reserved, including the making of copies unless permission is
obtained from the copyright proprietors. Pertinent discussion including author’s
(ACI 318-14) and 28.0 ~ 57.2 (Eurocode 2) for fy = 400 MPa closure, if any, will be published ten months from this journal’s date if the discussion
is received within four months of the paper’s print publication.

ACI Structural Journal/January 2018 235


Fig. 2—Comparison of lap-splice lengths by ACI 318-14 and Eurocode 2.
bond-splitting cracks in the lap-splice region become severe
owing to the presence of shear force, particularly after diag-
onal shear cracking. Repeated cyclic loading may also cause
bond deterioration in the plastic hinge zone. Ferguson and
Krishnaswamy12 evaluated the performance of lap splices
under moment gradient for bearing walls. They reported that
the performance of lap splices was affected by the average
stress of two spliced bars, which was less than the yield stress
because of the moment gradient of the lap-splice region.
They proposed a reduction factor of (1 + k)/2, ranging from
0.75 to 1.0, where k is the ratio of the less of two spliced
bar stresses to the yield stress. As no data for k < 0.5 were
available, the reduction factor for splice length was limited
to not less than 0.75.
In the previous studies, existing columns designed and
constructed before the 1970s were of concern because such
columns had relatively small lap-splice lengths of 20db and
30db, which are the minimum compression lap-splice lengths
Fig. 3—Types of lap splice using offset bars (ACI 315). specified in ACI 318-5613 and ACI 318-77,14 respectively. The
previous tests demonstrate that, even with such small lap-splice
splice, on the other hand, the top bars are offset inside. As lengths, the nominal flexural strength of the columns were
the bottom bars are located inside the core concrete, the attained, while the ductility and energy dissipation capacity
moment-carrying capacity of the column with the bottom were degraded owing to premature bond-splitting failure.
offset-bar splice is less. However, columns with the bottom In the present study, cyclic loading tests of columns with
offset-bar splice show better displacement ductility owing lap-spliced longitudinal bars in the plastic hinge region were
to the thicker concrete cover of the bottom offset-bar. For performed. The lap-splice length of the column bars, and the
this reason, in current seismic design codes, the bottom shear-span length (that is, moment gradient and shear force)
offset-bar splice is recommended (ACI 318-14, Section were taken into account as the primary test variables. With
10.7.5). On the contrary, if the top offset-bar splice is used, the results of the cyclic loading tests, the strength, ductility,
there is no reduction in the moment capacity due to the bar and failure mode of the columns were investigated and the
offset but the deformation capacity can be decreased due to effects of the test variables on the column performance were
premature bond-splitting cracks occurring along the splice discussed. Based on the results, the lap-splice length of
length (Boyes et al.10). Even in the case that the top and column bars was evaluated.
bottom bars are lap-spliced next to each other without offset,
the column performance such as the moment capacity and RESEARCH SIGNIFICANCE
deformation capacity is similar to that of columns with the Columns with lap splices in the plastic hinge region
top offset-bar splice.8 are permitted to use in ordinary and intermediate moment
The performance of lap splices is also affected by the frames. However, in most cases, effects of lap splices on the
distributions of flexural moment and shear force along the column behavior are not considered in seismic design and
length of columns.11,12 The lap-splice length of ACI 318-14 evaluation. In this study, the effects of moment gradient and
was developed considering uniform bond stresses based on shear force on the bond demand and splice performance were
the test results for beams with uniform moment (Fig. 4(a)). investigated. Based on the test results, recommendations
On the other hand, in columns subjected to lateral loading were made for the design and evaluation of these lap splices.
(Fig. 4(b)), the demand of bond stress varies along the The test results and design considerations could be useful in
splice length according to the moment gradient. However,

236 ACI Structural Journal/January 2018


Fig. 4—Bending moments and shear forces at lap-splice region.
the evaluation of existing buildings because old buildings or Table 1—Mechanical properties of steel
buildings in low and moderate seismic zones usually have reinforcement
short lap splices in the plastic hinge zones of columns. Test No.* Bar type db, mm fy, MPa fu, MPa
SD400 D13 12.7 500 645
TEST PLAN 1
Material strength SD500 D25 25.4 571 700
Eight column specimens were prepared for the cyclic load SD400 D13 12.7 481 607
tests. Table 1 presents the yield and tensile strengths (fy and fu, 2
SD500 D25 25.4 550 685
respectively) of the reinforcing steel bars used for the spec-
SD400 D13 12.7 528 656
imens. The yield strength and tensile strength are fy = 475 ~ 3
590 MPa (68.9 ~ 85.6 ksi) and fu = 607 ~ 707 MPa (88.0 ~ SD500 D25 25.4 588 702
102.5 ksi), respectively. In the table, SD400 and SD500 SD400 D13 12.7 475 659
denote the nominal yield strength grades of deformed bars, 4 SD500 D25 25.4 590 707
400 and 500 MPa (58.0 and 72.5 ksi), respectively, and D10,
SD500 D29 28.6 521 671
D13, D22, D25, and D29 denote the diameters of deformed
bars, db = 9.53, 12.7, 22.2, 25.4, and 28.6 mm ([0.375, 0.5, Eight columns were tested over four experiments. Thus, yield and tensile strength
*

differed even for reinforcing bars of same grade and size.


0.875, 1.0, and 1.125 in.), respectively.
Notes: 1 mm = 0.039 in.; 1 MPa = 0.145 ksi.
Table 2 presents the mixture proportion of the concrete.
The maximum aggregate size was 25.4 mm (1.0 in.). plastic hinge region with the bottom offset bars embedded
Compression tests of cylinders (100 x 200 mm [3.94 x in the base. Except for SL40S2B-1, the same diameter bars
7.87 in.]) were performed on the first day of testing. The were used for both the top and bottom bars. In SL40S2B-1,
compressive strength of the concrete was fc′ = 25 ~ 37 MPa the diameter of the bottom offset bars was increased to D29
(3.63 ~ 5.37 ksi). to make up for the moment strength reduction owing to the
bottom bar offset. The lap-splice lengths of the longitu-
Test specimens dinal bars varied from 30db to 50db, which were equivalent
The dimensions and reinforcement details of the eight to 57% ~ 108% of the required splice lengths of ACI 318-14
square column specimens are shown in Fig. 5. The test vari- calculated using the actual material strength (fc′ and fy) and
ables such as the shear span-to-height ratio (a/h), lap-splice bottom offset-bar diameters (db).
length (ls), and transverse hoop spacing (s) are presented in The lap-spliced longitudinal bars were confined by trans-
Table 3. In the specimen names, the first letter “S” denotes verse hoops with 90-degree end hooks for anchorage. SD400
square cross section; the following letters “L00”, “L30”, D13 hoops were used at a spacing of s = 165 mm (6.5 in.)
“L40”, and “L50” denote the length of lap splice, ls = 00db (= 0.5d; d is the effective depth of the column), except for
(that is, without lap splice), 30db, 40db, and 50db, respec- SL50S2B-1 with s = 330 mm (13.0 in.) (=1.0d). The first hoop
tively; “S1” and “S2” denote the shear spans, a = 1200 and was placed at a height of 80 mm (3.15 in.) from the base. As
2400 mm (47.2 and 94.5 in.), respectively; and the last letter shown in Fig. 5, the location of the 90-degree hook anchorage
“B” denotes the offset details of column bars (that is, bottom of the hoops was alternated in accordance with ACI 318-14.
offset bar splice, refer to Fig. 3(a)).
The cross-sectional dimensions of the columns were 400 x Loading method
400 mm (15.8 × 15.8 in.) (h = 400 mm [15.8 in.]). The shear Figure 6 shows the test setup for lateral and axial loading.
span lengths were a = 1200 mm (47.2 in.) (a/h = 3.0) or The cyclic lateral loading was controlled by the lateral
a = 2400 mm (94.5 in.) (a/h = 6.0). SD500 D25 and D29 displacement of the actuator with a maximum stroke of
bars were used for the longitudinal reinforcement. In all ±250 mm (9.84 in.). The actuator was placed at 1200 or
specimens, the straight top bars were lap-spliced at the 2400 mm (47.2 or 94.5 in.) from the column base. During

ACI Structural Journal/January 2018 237


Table 2—Mixture design of concrete
Unit weight, kgf/m3
Test No. Compressive strength fc′ w/c W C FS S FA CA SP Slump
1 32 MPa 55.5% 155 153 56 70 923 950 1.95 120 mm
2 25 MPa 46.2% 173 243 56 75 807 887 2.24 150 mm
3 27 MPa 53.5% 159 149 74 74 907 893 1.78 120 mm
4 37 MPa 44.8% 168 319 56 - 832 915 2.25 150 mm
Notes: W is water; C is cement; FS is fly ash; S is blast-furnace slag; FA is fine aggregate; CA is coarse aggregate; SP is superplasticizer; 1 mm = 0.039 in.

Table 3—Summary of test variables


Specimen Test No. Shear span length a Splice length ls Hoop spacing s Number of hoops along lap splice Ratio of bottom splice offset bars
SL00S1B 1 N/A 0.5d — —
1200 mm
SL30S1B 1 30db 0.5d 4
(a/h = 3.0)
SL40S1B 1 40db 0.5d 6
SL30S2B 2 30db 0.5d 4 2.53% (8-D25)
SL40S2B 2 40db 0.5d 6
2400 mm
SL50S2B 4 50db 0.5d 8
(a/h = 6.0)
SL40S2B-1 4 40db 0.5d 7 3.21% (8-D29)
SL50S2B-1 3 50db 1.0d 4 2.53% (8-D25)
Note: 1 mm = 0.039 in.

Fig. 5—Dimensions and reinforcement details of specimens. (Note: 1 mm = 0.039 in.)


cyclic loading tests, the lateral drift ratio of the column was An axial load N (= 0.15 to 0.185Agfc′) was applied to the
increased to 0.25%, 0.35%, 0.5%, 0.75%, 1.0%, 1.5%, 2.0%, top of the specimen by post-tensioning two high-strength
2.5%, 3.5%, 5.0%, and 7.0%. The loading sequence was bars (diameter = 47 mm [1.85 in.]). For post-tensioning,
planned in accordance with ACI 374.1.15 Thus, the target two hydraulic jacks with a capacity of 1000 kN (225 kip)
lateral drift ratio of a loading step was determined as 1.25 ~ were placed on the top of the column. During cyclic lateral
1.5 times the previous loading step, and load cycles were loading, the level of axial load was kept uniform by main-
repeated three times at each loading step. The loading rate taining the hydraulic pressure of the jacks. As shown in
was increased from 20 mm/min (0.787 in/min) for δ = 0.25% Fig. 6, because pin joints are provided at the bottom of the
to 60 mm/min (2.36 in./min) for δ = 1.0%, and then main- bars, the bars are free to move laterally. Because the axial
tained a constant value of 60 mm/min (2.36 in./min) after load of the bars always acts along the centerline of the
δ = 1.0%. column, the second-order effect is negligible in the lateral
load-displacement relationship.

238 ACI Structural Journal/January 2018


Fig. 6—Test setup for axial and lateral loading. (Note: 1 mm = 0.039 in.)
TEST RESULTS bottom of the column (Fig. 8(b)). In SL40S1B, with the
Cyclic behavior and failure mode increased splice length ls = 40db, the maximum loads
Figure 7 shows the base moment and lateral drift ratio (Mu = +393 and –417  kN∙m [+290 and –308 kip∙ft]) were
(M-δ) relationships of the columns. For a fair strength significantly increased when compared to SL30S1B. Despite
comparison between the columns with different shear spans, the lower nominal strength, the test strength was equivalent
the base moment M (= Pa, where P is the lateral load of to that of SL00S1B without lap splice. The enhanced strength
the actuator and a is the shear span of the column) is used, of SL40S1B is attributed to the increased splice length.
instead of the lateral load P. The lateral drift ratio δ is defined Spalling of the cover concrete began to occur at δ = 2.0%.
by dividing the lateral displacement at the loading point by The load-carrying capacity was significantly degraded at the
the shear span. The maximum test moments Mu are denoted first load cycle of δ = –5.0% owing to excessive concrete
as white circles, while the calculated moment strength Mn crushing and web shear cracking (Fig. 8(c)).
without strength reduction factor is denoted as horizontal Figures 7(d) and (e) show the cyclic behaviors of
dashed lines. The Mn of each specimen was calculated SL30S2B and SL40S2B with the increased shear span a =
from section analysis, considering the applied axial load N, 2400 mm (94.5 in.). The maximum test loads were approx-
measured material strength, and location of the bottom offset imately 95% of the nominal strength (Mn = 294 kN∙m
bars in the cross section. [217 kip∙ft]), which indicates that the performance of the lap
In SL00S1B, without lap splices, the maximum loads splices was not successful. Despite the insufficient strength,
(Mu  = +407 and –400 kN∙m [+300 and –295 kip∙ft]) their ductile behaviors were maintained until δ = 7.0%. In
were greater than the calculated moment strength (Mn = SL30S2B and SL40S2B, the column shear was significantly
±374 kN∙m [±276 kip∙ft]). Strength degradation occurred at reduced because of the increased shear-span length, when
δ = 2.0% after spalling of the concrete cover at the bottom compared to those of SL30S1B and SL40S1B with a shorter
of the column. Ultimately, at δ = 3.5%, the load-carrying shear span. Thus, as shown in Fig. 8(d) and (e), diagonal
capacity was significantly decreased because of hoop failure shear cracking in the web was not significant even at large
at the 90-degree hook anchorages (Fig. 8(a)). Despite the inelastic deformations of δ = 5.0% or greater, and conse-
presence of the axial load of N = 870 kN (196 kip), severe quently the displacement capacity was increased. However,
web shear cracking occurred. Concrete strut failure occurred bond-splitting cracks occurred owing to the insufficient
in the plastic hinge region. splice lengths (that is, 57% and 76% of the required splice
In SL30S1B, with lap splice ls = 30db (63% of the minimum length) and propagated along the spliced longitudinal bars
splice length required by ACI 318-14), the maximum loads as the inelastic deformation increased. When the test results
(Mu= +350 and –354 kN∙m [+258 and –261 kip∙ft]) were of SL30S2B and SL40S2B are compared with those of
13% less than those of SL00S1B without lap splice. Never- SL30S1B and SL40S1B, obviously, the shear span length a
theless, Mu was slightly greater than the moment strength affected the failure mode and ductility of the columns with lap
(Mn = ±340 kN∙m [251 kip∙ft]). Because the bottom bars, splice: when the shear-span ratio was decreased (a/h = 3.0),
which resist flexure moment, were offset inside (Fig. 3(a)), even with the short splice length, the test strength exceeded
the test strength and nominal flexural strength of SL30S1B the nominal flexural strength. On the other hand, when the
were less than those of SL00S1B. Ultimately, failure shear span ratio was increased (a/h = 6.0), the test strength
occurred at δ = 3.5% because of concrete crushing at the was less than the nominal flexural strength. However, despite

ACI Structural Journal/January 2018 239


Fig. 7—Base moment (M)-lateral drift ratio (δ) relationships of columns. (Note: 1 in. = 25.4 mm; 1 kip = 4.448 kN.)

Fig. 8—Cracking modes at δ = 2.0% and failure modes at end of tests.

240 ACI Structural Journal/January 2018


Table 4—Summary of test results
Lap-splice length (mm or ratio)
Specimen fc′, MPa ls ls/db ls,ACI* ls/ls,ACI lsm† ls/lsm Mu‡, kN∙m Mn, kN∙m Mu/Mn Δy§, mm Δu||, mm Failure mode#
SL00S1B 32 — — — — — — 404 374 1.08 9.3 40 FS
SL30S1B 32 750 30 1200 0.63 936 0.80 352 340 1.04 7.1 44 FS
SL40S1B 32 1000 40 1200 0.83 936 1.07 405 340 1.19 10.1 69 FS
SL30S2B 25 750 30 1308 0.57 1240 0.60 278 294 0.94 18.5 132 BS
SL40S2B 25 1000 40 1308 0.76 1240 0.81 281 294 0.95 17.9 168 —
SL50S2B 37 1250 50 1153 1.08 1266 0.99 371 367 1.01 23.8 168 —
SL40S2B-1 37 1160 40 1146 1.01 1128 1.11 363 393 0.92 22.8 168 —
SL50S2B-1 27 1250 50 1345 0.93 1196 0.97 342 312 1.10 22.2 120 BS
*
Length of Class B lap splice specified in ACI 318-14 on basis of bottom offset bars.

Modified lap-splice length in Eq. (3).

Average of positive and negative maximum loads.
§
Δy = Pu/Ky, where Pu is maximum load and Ky is secant stiffness connecting origin and pre-peak point of 0.6Pu.
||
Δu is defined as average value of positive and negative maximum displacement corresponding to 80% of maximum load or as displacement at end of tests.
#
FS and BS, respectively, denote web shear failure after flexural yielding, bond-splitting failure along lap-splice length. In SL40S2B, SL50S2B, and SL40S2B-1, tests were termi-
nated without failure because of limited actuator stroke.
Notes: 1 in. = 25.4 mm; 1 kip = 4.448 kN; 1 ksi = 6.895 MPa.

the lower strength, the ductile behaviors of SL30S2B and ductility and energy dissipation capacity of SL50S2B-1 were
SL40S2B were maintained until large inelastic deformations. significantly degraded.
In SL50S2B with ls = 50db, the maximum load (Mu =
+364 and –377 kN∙m [268 and 278 kip∙ft]) was signifi- Strains of lap-spliced bars
cantly increased, reaching to the nominal strength (Mn = To investigate the force transfer between the spliced
±367  kN∙m [±271 kip∙ft]). Because the splice length was longitudinal bars, the strains of the lap-spliced bars were
greater than the requirement of ACI 318-14, the strength measured using strain gauges. Figure 9 shows the spliced
degradation during repeated load cycles was limited and the bar strains measured from SL30S2B, SL40S2B, SL50S2B,
energy dissipation capacity, represented as the enclosed area SL40S2B-1, and SL50S2B-1. In each specimen, the strains
by a full load cycle in the load-displacement curves shown in of the straight bar and offset bar are plotted in the right and
Fig. 7, was significantly enhanced when compared to other left figures, respectively. The horizontal and vertical axes,
columns with ls = 30db and 40db. At δ = 7.0%, the test was respectively, denote the bar strains and height from the base
terminated because of lack of actuator stroke. As shown in where the bar strains were measured. The bar strains in
Fig. 8(f), bond-splitting cracking was not significant, which Fig. 9 were the maximum tensile strains that each spliced bar
indicates that the force transfer through the bond between experienced during the load cycles repeated at each lateral
the concrete and spliced bars was successful. drift ratio. The bar strains corresponding to δ = 0.25%, 0.5%,
In SL40S2B-1 with bottom offset bars of a greater bar diam- 1.0%, 1.5%, 2.0%, and 2.5% are presented.
eter D29 (ls = 40db and db = 28.6 mm [1.125 in.]), the maximum In Fig. 9, at each lateral drift ratio, the strains of the
loads (Mu = +342 and –383 kN∙m [252 and 282 kip∙ft]) were bottom offset bars decreased as the height from the base
less than the calculated moment strength (Mn = ±393 kN∙m increased, while the strains of the straight top bars did not
[±290 kip∙ft]) using the measured material strength (Fig. 7(g)). significantly vary along the lap-splice length. Further, the
This indicates that the lap-splice length ls = 40db was not suffi- strains of the bottom offset bars were much greater than
cient to develop the full yield strength of the D29 offset bars, those of the straight top bars. This result indicates that owing
although ls/ls,ACI was greater than 1.0 (Table 4). However, to the moment gradient along the lap-splice length, the bond
as the spliced bars were confined by transverse hoops with demand of the bottom offset bars was much greater than that
a spacing s = 0.5d, the strength degradation was limited and of the straight top bars.
the hysteretic energy dissipation was significant. On the The magnitude of the bar strains was significantly
other hand, in SL50S2B-1 with the increased hoop spacing affected by the lap-splice length. The bar strains of SL30S2B
s = 1.0d, the maximum loads (Mu = +347 and –337  kN∙m (ls/ls,ACI = 0.57) and SL40S2B (ls/ls,ACI = 0.76) in Fig. 9(a) and
[256 and 249 kip∙ft]) were greater than the calculated moment (b) were smaller than the yield strain (εy = 0.00275). Because
strength (Mn = ±312 kN∙m [±230 kip∙ft]), which indicates that of the short splice lengths, the stress of the bottom offset bars
the lap-splice length ls = 50db satisfactorily developed the full was not fully transferred to the straight top bars. The bar
yield strength of the D25 offset bars. However, as the hoop strains of SL40S2B with greater splice length were greater
spacing s = 1.0d was not sufficient to ensure a post-yield than those of SL30S2B with a less splice length.
ductile behavior of the column, bond-splitting cracks occurred In SL50S2B (ls/ls,ACI = 1.08) and SL50S2B-1 (ls/ls,ACI =
along the lap splice at δ = 2.0% (Fig. 8(h)). Consequently, the 0.93), on the other hand, the bar strains were significantly
increased by increasing the lap-splice lengths. The strains

ACI Structural Journal/January 2018 241


Fig. 9—Strains of spliced bars at lateral drift ratios 0.25% ~ 2.5%. (Note: 1 mm = 0.039 in.)
tensile stress fsO of the bottom offset bars and the tensile
stress fsS of the straight top bars. As shown in Fig. 10, in
the case of SL30S2B and SL40S2B (a/h = 6.0), the moment
Ms at the top end of the lap-splice region is greater, thereby
increasing stress fsS of the straight top bars and the average
bond demand along the lap-splice length (=[fsO + fsS]/2).
On the other hand, in SL30S1B and SL40S1B (a/h = 3.0),
the average bond demand is smaller. For this reason, the
nominal flexural strength was developed despite the short
splice length (ls/ls,ACI = 0.63 ~ 0.83).

Transverse reinforcement
In SL30S2B and SL40S2B with a hoop spacing of s =
Fig. 10—Moment gradient and spliced bar stresses. (Note: 0.5d, despite the short splice length, ductile behavior was
1 mm = 0.039 in.) maintained, although the test strength was less than the
nominal flexural strength. On the other hand, in SL50S2B-1
of the bottom offset bars exceeded the yield strains (εy =
with the increased hoop spacing of s = 1.0d, web shear
0.00263 or 0.00295). The strain distribution of SL40S2B-1
cracking became severe and vertical bond-splitting cracks
(ls/ls,ACI = 1.01) was similar, although the size of the bottom
propagated along the spliced bars, thereby decreasing the
offset bars was increased to D29 (db = 28.6 mm [1.125 in.]).
deformation capacity. This result indicates that closely
spaced hoops should be used to maintain ductility, avoiding
FACTORS AFFECTING PERFORMANCE OF
premature bond failure under cyclic loading. As illustrated
LAP SPLICE
in Fig. 10, the confining hoops can provide a clamping force
Moment gradient and lap-splice length
across the splitting crack.
As shown in Fig. 7 and 8, the performance of the columns
with lap splices at the plastic hinge was significantly affected
LAP-SPLICE LENGTH AFFECTED BY
by the shear span length. In SL30S1B and SL40S1B with a
MOMENT GRADIENT
smaller shear span length, a = 1200 mm (47.2 in.) (a/h = 3)
In columns, because the bond demand of the spliced bars
(refer to Fig. 7(b) and (c)), the maximum loads Mu were
varies with the moment gradient, the lap-splice length can
greater than the moment strength Mn, although the lap-splice
be redefined considering the moment gradient. The bond
lengths ls were only 63% or 83% of minimum splice lengths
demand for a lap-splice length can be defined as the average
ls,ACI specified in ACI 318-14. In contrast, in SL30S2B and
of the tensile stresses of the bottom offset bars fsO and
SL40S2B with a greater shear span length, a = 2400 mm
straight top bars fsS. fsO can be taken as the yield strength
(94.5 in.) (a/h = 6) (refer to Fig. 7(d) and (e)), the maximum
fyO of the bottom offset bars. On the other hand, fsS of the
loads Mu were less than the moment strength Mn. The lower
straight column bars, which is less than the yield strength
strength of SL30S2B and SL40S2B is attributed to the
fyS, can be estimated considering the moment gradient as
increased bond demand of the spliced bars, which involved
follows (refer to Fig. 10)
the moment gradient in the lap-splice region.
Figure 10 shows the effect of the moment gradient in the
lap slice region. The tensile stress gradients of the bottom  M   l 
f sS ≈ 1 − s  f yO = 1 − s  f yO ≤ f yS (1)
offset bars and straight top bars along the lap-splice length  Mn   a
are also illustrated in the same figure. The solid and dotted
where Mn is the nominal flexure strength at the bottom of
lines denote the moment gradients and spliced bar stresses
the column; Ms is the moment demand at the top of the lap
corresponding to a = 1200 and 2400 mm (47.2 and 94.5 in.)
splice (Mnls/a); a is the shear span length of the column; and
(a/h = 3.0 and 6.0), respectively. The bond demand over
ls is the length of the lap splice. It is noted that, in Eq. (1), fsS
the entire lap-splice length can be defined as the sum of the
of the straight column bars does not depend on fyS, but on fyO

242 ACI Structural Journal/January 2018


Fig. 11—Lap-splice and offset details of existing column specimens.
of the bottom offset bars considering yielding at the bottom is the modified lap-splice length calculated by Eq. (3). In
of the column. general, the Mu/Mn ratio tends to increase as the lst/lsm ratio
By replacing fy in ACI 318-14’s Class B lap-splice length increases. In all columns with lst/lsm ratios not less than 1.0,
equation with the average bond demand (fyO + fsS)/2 = fyO except SL40S2B-1, which has different bar diameters in the
(1 – lsm/[2a]), the modified lap-splice length lsm in columns spliced bars, the Mu/Mn ratios were equal to or greater than
can be obtained as follows 1.0. This result indicates that Eq. (3) safely estimated the
lap-splice lengths required to develop the nominal flexural
 l  strength in columns with moment gradient. Furthermore, as
lsm = 1.3 1 − sm  ldO (2) shown in Fig. 12(b), the ductility of the test specimens tends
 2a 
to increase when the lst/lsm ratio increases. The displacement
ductility (Δu/Δy) exceeds 5.0 when the lst/lsm ratio is greater
 2  than approximately 1.2. Δu and Δy are the ultimate and yield
or lsm =  ldO (3)
 1.54 + ldO a  displacements presented in previous studies, respectively.
where ldO is the basic development length of the bottom SUMMARY AND CONCLUSIONS
offset bars specified in ACI 318-14. In Eq. (3), the range of Cyclic loading tests of columns with lap splices in the
the modified lap-splice length is defined as 0.87ldO ≤ lsm ≤ plastic hinge region were performed. Based on the test results,
1.3ldO depending on the moment gradient. When shear span the effects of lap-splice length, moment gradient, and trans-
a is equal to 1.3ldO, Eq. (3) yields the minimum value of the verse reinforcement on the failure mode and deformation
lap-splice length (lsm = 0.87ldO). In the case of shear spans capacity were investigated. Further, a modified lap-splice
a far greater than ldO, Eq. (3) yields the maximum value length for columns with moment gradient was proposed. The
of lsm = 1.3ldO. For conservative design and evaluation, the major findings of this study are summarized as follows.
shear span length a may be taken as the clear height between 1. In the specimens with lap-splice length (ls = 30db or
the top and bottom floors. The proposed method is similar to 40db) less than the requirement of ACI 318-14, when the
the result of Ferguson and Krishnaswamy.12 shear span ratio was a/h = 3.0 (SL30S1B and SL40S1B),
For verification, the modified splice length was compared the test strength reached the nominal flexural strength,
with the test results of the present and existing studies. despite the short splice length. On the other hand, when the
Figure 11 shows the lap-splice and offset details used in shear-span ratio was a/h = 6.0 (SL30S2B and SL40S2B),
the existing column specimens considered for the verifi- the test strength was less than the nominal strength (Mu =
cation. The test results of the present and existing studies 0.95Mn); however, ductile behavior was maintained until
are summarized in Table 4 and Table 5, respectively. Twen- large inelastic deformation.
ty-six columns were collected from 10 studies.3-8,10,16-18 2. In Specimen SL50S2B (ls = 50db), with the lap-splice
The ranges of test parameters are 19.7 MPa (2.86 ksi) ≤ fc′ length satisfying the requirement of ACI 318, despite the
≤ 41.4 MPa (6.0 ksi), 315 MPa (45.7 ksi) ≤ fy ≤ 617 MPa lap splice being located in the plastic hinge zone, the test
(89.5 ksi), 12.5 mm (0.49 in.) ≤ db ≤ 31.8 mm (1.25 in.), 105 strength reached the nominal flexural strength and the ductile
mm (4.13 in., 0.21Hmin) ≤ s ≤ 457.2 mm (18.0 in., 1.0Hmin), behavior occurred. However, in Specimen SL50S2B-1,
and 280 mm (11.0 in., 20db) ≤ ls ≤ 880 mm (34.6 in., 40db). with a greater spacing of hoops (s = 1.0d), the deformation
Figure 12(a) shows the relationships between the capacity was decreased.
maximum test load-to-nominal strength ratio (Mu/Mn) and Based on the present and previous test results, the
the provided-to-predicted lap-splice length ratio (lst/lsm), following considerations for the design and evaluation of
where lst is the lap-splice length used in the tests, and lsm

ACI Structural Journal/January 2018 243


Table 5—Summary of existing tests of columns with lap splices
Specimen properties Lap-splice length Moment strength
Specimen h, mm fc′, MPa fy, MPa N/Agfc′ a, mm s/h db, mm ls, mm lsm, mm ls/lsm Mu, kN·m Mn, kN·m Mu/Mn
3SLH18 25.6 331 0.09 2946 1.0 31.8 635 1141 0.56 415 449 0.92
Lynn et al. 3
2SLH18 457 33.1 331 0.07 2946 1.0 25.4 508 772 0.66 454 333 1.36
3SMD12 25.5 331 0.28 2946 0.67 31.8 635 1091 0.58 590 505 1.17

Aboutaha FC4 19.7 414 0 2743 0.89 25.4 610 1141 0.53 488 553 0.88
457
et al.4 FC5 19.7 414 0 2743 0.89 25.4 610 1141 0.53 488 553 0.88
2S10M 36 510 0.1 1829 0.67 25.4 508 966 0.53 371 432 0.86
2S20M 36 510 0.2 1829 0.67 25.4 508 966 0.53 427 492 0.87

Melek and 2S30M 36 510 0.3 1829 0.67 25.4 508 966 0.53 522 514 1.02
457
Wallace5 2S20H 35 510 0.2 1676 0.67 25.4 508 953 0.53 452 492 0.92
2S20HN 35 510 0.2 1676 0.67 25.4 508 953 0.53 448 492 0.91
2S30X 35 510 0.3 1524 0.67 25.4 508 926 0.55 519 514 1.01
Haroun and
RF-A1 610 41.4 443 0.054 3429 0.21 19.0 381 895 0.43 0.832 1089 0.76
Elsanadedy6
Harries et al.7 L0 458 27.6 414 0.26 2440 0.78 22.2 490 873 0.56 426 389 1.09
C14 39 550 0 1400 0.5 1.0 420 591 0.71 111 119 0.93
Harajili8 C16 400 40 528 0 1400 0.5 16.0 480 629 0.76 135 132 1.02
C20 32 617 0 1400 0.5 20.0 600 1040 0.58 122 183 0.67
24L-300-2D 33.6 315 0.3 1624 0.67 25.4 600 667 0.90 406 406 1.00
Boyes et al.10 450
30L-300-2D 33.9 315 0.3 1624 0.67 25.4 750 664 1.13 426 406 1.05

Bournas L20d_C 27.8 523 0.28 1600 0.8 14.0 280 668 0.42 62 67 0.93
250
et al.16 L40d_C 25.8 523 0.28 1600 0.8 14.0 560 688 0.81 72 67 1.08

ElGawady AB-1 31 331 0.07 1803 0.33 12.5 445 401 1.11 135 112 1.21
381
et al.17 AB-2 31 331 0.07 1803 0.33 12.5 445 401 1.11 130 112 1.16
RL30S1T 32 566 0.17 1200 0.42 22.2 660 859 0.77 227 208 1.09
RL40S1T 32 566 0.10 1200 0.42 22.2 880 859 1.02 206 202 1.02
Kim et al.18 400
RL30S2T 25 565 0.175 2400 0.42 22.2 660 1149 0.57 191 183 1.04
RL40S2T 25 565 0 2400 0.42 22.2 880 1149 0.77 162 164 0.99
Notes: 1 mm = 0.039 in.; 1 MPa = 0.145 ksi.

Fig. 12—Moment strength ratio and ductility according to lap-splice length ratio.
the splice length are recommended for ordinary and interme- 2. In the columns with offset longitudinal bars for lap
diate moment frames. splice, because the flexural strength can be decreased owing
1. When the lap-splice length satisfies the requirements of to the location of the bottom offset bars, the nominal flexural
ACI 318-14, a lap splice of longitudinal bars can be used in strength Mn should be calculated considering the locations of
the plastic hinge zone of columns. the offset bars in the cross section.

244 ACI Structural Journal/January 2018


3. To avoid premature bond failure, lap-spliced longitu- α4 = influencing factor for confinement by welded transverse
reinforcement
dinal bars should be confined by closely spaced transverse α5 = influencing factor for confinement by transverse pressure
reinforcement. In this study, s = 0.5d (or 165 mm [6.50 in.]) α6 = influencing factor for lapped bars relative to total cross section
is recommended for Grade 500 MPa (72.5 ksi) D25 longi- area
δ = drift ratio (=Δ/a)
tudinal reinforcing bars (bar diameter = 25.4 mm [1.0 in.]); εy = yield strain of steel bar
Otherwise, the deformation capacity and energy dissipation μ = ductility (=Δu/Δy)
are significantly decreased, even with the splice length spec-
ified in ACI 318-14. REFERENCES
1. ACI Committee 318, “Building Code Requirements for Structural
Concrete (ACI 318-14) and Commentary (ACI 318R-14),” American
AUTHOR BIOS Concrete Institute, Farmington Hills, MI, 2014, 519 pp.
Chul-Goo Kim is a Researcher in the Institute of Engineering Research at
2. Eurocode 2, “Design of Concrete Structures: Part 1-1: General Rules
Seoul National University, Seoul, South Korea, where he received his BE,
and Rules for Buildings,” British Standards Institution, London, UK, 2004.
MS, and PhD in architectural engineering. His research interests include
3. Lynn, A. C.; Moehle, J. P.; Mahin, S. A.; and Holmes, W. T., “Seismic
shear behavior and earthquake design of reinforced concrete structures.
Evaluation of Existing Reinforced Concrete Building Columns,” Earth-
quake Spectra, V. 12, No. 4, 1996, pp. 715-739. doi: 10.1193/1.1585907
ACI member Hong-Gun Park is a Professor in the Department of Architec-
4. Aboutaha, R. S.; Engelhardt, M. D.; Jirsa, J. O.; and Kreger, M. E.,
ture & Architectural Engineering at Seoul National University. He received
“Experimental Investigation of Seismic Repair of Lap Splice Failures
his BE and MS in architectural engineering from Seoul National University
in Damaged Concrete Columns,” ACI Structural Journal, V. 96, No. 2,
and his PhD in civil engineering from the University of Texas at Austin,
Mar.-Apr. 1999, pp. 297-306.
Austin, TX. His research interests include numerical analysis and seismic
5. Melek, M., and Wallace, J. W., “Cyclic Behavior of Columns with
design of reinforced concrete and composite structures.
Short Lap Splices,” ACI Structural Journal, V. 101, No. 6, Nov.-Dec. 2004,
pp. 802-811.
Tae-Sung Eom is an Associate Professor in the Department of Architec-
6. Haroun, M. A., and Elsanadedy, H. M., “Fiber-Reinforced Plastic Jackets
tural Engineering at Dankook University, Gyeonggi-do, South Korea. He
for Ductility Enhancement of Reinforced Concrete Bridge Columns with
received his BE, MS, and PhD in architectural engineering from Seoul
Poor Lap-Splice Detailing,” Journal of Bridge Engineering, ASCE, V. 10,
National University. His research interests include experiment and numer-
No. 6, 2005, pp. 749-757. doi: 10.1061/(ASCE)1084-0702(2005)10:6(749)
ical analysis of reinforced concrete and composite structures.
7. Harries, K. A.; Ricles, J. M.; Pessiki, S.; and Sause, R., “Seismic
Retrofit of Lap Splices in Nonductile Square Columns Using Carbon
ACKNOWLEDGMENTS Fiber-Reinforced Jackets,” ACI Structural Journal, V. 103, No. 6, Nov.-Dec.
This research was supported by a grant (Code 15AUDP-B066083-03) 2006, pp. 874-884.
from the R&D Policy Infra program funded by the Ministry of Land, Infra- 8. Harajli, M., “Seismic Behavior of RC Columns with Bond-Critical
structure, and Transportation (MOLIT) of the Korean government, and the Regions: Criteria for Bond Strengthening Using External FRP Jackets,”
research fund of National Research Foundation of Korea (Code R-2015- Journal of Composites for Construction, ASCE, V. 12, No. 1, 2008,
00441, Mid-career Research Program). The Institute of Engineering pp. 69-79. doi: 10.1061/(ASCE)1090-0268(2008)12:1(69)
Research at Seoul National University also provided research facilities for 9. ACI Committee 315, “Details and Detailing of Concrete Reinforce-
this work. The authors are grateful for their support. ment (ACI 315-99),” American Concrete Institute, Farmington Hills, MI,
1999, pp. 1-44.
10. Boyes, A.; Bull, D. K.; and Pampanin, S., “Seismic Performance of
NOTATION Concrete Columns with Inadequate Transverse Reinforcement,” University
cb = lesser of: (a) distance from center of bar to nearest concrete
of Canterbury, Christchurch, New Zealand, 2008, 12 pp.
surface; and (b) one-half center-to-center spacing of bars being
11. Paulay, T., “Lapped Splices in Earthquake-Resisting Columns,” ACI
developed
Journal Proceedings, V. 79, No. 6, June 1982, pp. 458-469.
d = distance from extreme compression fiber to centroid of longitu-
12. Ferguson, P. M., and Krishnaswamy, C., “Tensile Lap Splices-Part
dinal tension reinforcement
2: Design Recommendations for Retaining Wall Splices and Large Bar
db = nominal diameter of bar
Splices,” Research Report No. 113-3, Center for Highway Research,
fbd = concrete bond strength
University of Texas at Austin, Austin, TX, July 1971, 67 pp.
fc′ = specified compressive strength of concrete
13. ACI Committee 318, “Building Code Requirements for Reinforced
fck = characteristic compressive cylinder strength of concrete at 28 days
Concrete (ACI 318-56),” American Concrete Institute, Farmington Hills,
fsO = tensile stress of bottom offset bar
MI, 1956, pp. 1-73.
fsS = tensile stress of straight top bar
14. ACI Committee 318, “Building Code Requirements for Reinforced
fu = ultimate strength for nonprestressed reinforcement
Concrete (ACI 318-77),” American Concrete Institute, Farmington Hills,
fy = specified yield strength for nonprestressed reinforcement
MI, 1977, pp. 1-103.
fyd = design value of yield strength of steel
15. ACI Committee 374, “Acceptance Criteria for Moment Frames
fyO = yield stress of bottom offset bar
Based on Structural Testing and Commentary (ACI 374.1-05),” American
fyS = yield stress of straight top bar
Concrete Institute, Farmington Hills, MI, 2005, pp. 1-9.
hmin = minimum column dimension
16. Bournas, D. A.; Triantafillou, T. C.; Zygouris, K.; and Stavropoulos, F.,
Ktr = transverse reinforcement index
“Textile-Reinforced Mortar Versus Frp Jacketing in Seismic Retrofitting of
ldO = basic development length of bottom offset bars
RC Columns with Continuous or Lap-Spliced Deformed Bars,” Journal of
ls = lap splice length
Composites for Construction, ASCE, V. 13, No. 5, 2009, pp. 360-371. doi:
lsm = modified lap splice length
10.1061/(ASCE)CC.1943-5614.0000028
Mn = nominal flexural strength at section
17. ElGawady, M.; Endeshaw, M.; McLean, D.; and Sack, R., “Retro-
Ms = moment demand at top of lap splice
fitting of Rectangular Columns with Deficient Lap Splices,” Journal of
Mu = maximum test moment
Composites for Construction, ASCE, V. 14, No. 1, 2010, pp. 22-35. doi:
s = center-to-center spacing of transverse reinforcement
10.1061/(ASCE)CC.1943-5614.0000047
α1 = influencing factor for shape of bars
18. Kim, C. G., Park, H. G., Eom, T. S., and Kim, T. W., “Performance of
α2 = influencing factor for concrete minimum cover
Lap Splices in Rectangular RC Columns,” Proceedings of Korea Institute
α3 = influencing factor for confinement by transverse reinforcement
for Structural Maintenance and Inspection, Oct. 2015, pp. 251-254.
not welded to main reinforcement

ACI Structural Journal/January 2018 245


NEW
Symposium
Publications from ACI

SP-304, Sustainable Performance SP-305, Durability and Sustainability


of Concrete Bridges and Elements of Concrete Structures
Subjected to Aggressive In October 2015, the Italy Chapter – ACI (ACI IC) and
Environments: Monitoring, the Department of Civil, Chemical, Environmental, and
Material Engineering (DICAM) of the University of
Evaluation, and Rehabilitation Bologna sponsored the First International Workshop
This collection contains 10 papers selected from on “Durability & Sustainability of Concrete Structures”
the three special sessions held at The Concrete in Bologna, Italy. The proceedings of the workshop
Convention and Exposition in Washington, DC, consist of 48 papers concerning reduction in
October 2014. Topics include: Performance greenhouse gases in cement and concrete industry,
Reliability of Reinforced Concrete Beams recycled materials, innovative binders and
Strengthened with Fiber-Reinforced Polymer geopolymers, life-cycle cost assessment in concrete
(FRP) Sheets, Reducing Deck Cracking in construction, reuse and functional resilience of
Composite Bridges by Controlling Long-Term reinforced concrete structures, repair and main-
Properties, and many more. tenance, testing, inspection, and monitoring.

Available in CD/DVD or PDF format: $76.50 Available in PDF format: $101.50


(ACI members: $46.00) ($30.50 savings) (ACI members: $61.00) ($40.50 savings)

+1.248.848.3700 • www.concrete.org    
ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 115-S20

Conventional and High-Strength Steel Hooked Bars:


Detailing Effects
by J. Sperry, D. Darwin, M. O’Reilly, A. Lepage, R. D. Lequesne, A. Matamoros, L. R. Feldman,
S. Yasso, N. Searle, M. DeRubeis, and A. Ajaam
Findings from a study on the effect of hook bend angle, concrete confining reinforcement on hook anchorage strength for a
clear cover, and orientation of confining reinforcement on hook broader range of steel and concrete strengths than used in the
anchorage strength are presented. The range of test parameters earlier studies. Additional results and analyses are presented
was much broader than in previous studies. Bar stress at anchorage by Sperry et al. (2015a,b; 2017a,b).
failure ranged from 33,000 to 137,400 psi (228 to 947 MPa) and
concrete compressive strengths ranged from 4300 to 16,500 psi
RESEARCH SIGNIFICANCE
(30 to 114 MPa). Anchorage strength of hooked bars was insensi-
tive to bend angle (90 or 180 degrees) and side cover (between 2.5 The use of high-strength steel and concrete as a means
and 3.5 in. [65 and 90 mm]). Confining reinforcement was found to of reducing reinforcement congestion, member dimensions,
increase anchorage strength for 180-degree hooked bars regardless and material use has increased. The analysis reported herein
of orientation (parallel or perpendicular to the embedment length). is the first to evaluate whether previous findings showing
For 90-degree hooked bars, reinforcement oriented parallel to the that bars with 90- and 180-degree hooks have equivalent
embedment length had a greater effect on anchorage strength than anchorage strengths are valid for high-strength materials
reinforcement oriented perpendicular to the embedment length. and the first to establish the effect of increasing concrete
clear cover from 2.5 to 3.5 in. (65 to 90 mm). The effective-
Keywords: anchorage; beam-column joints; bond; development length;
high-strength concrete; high-strength steel; hook bend angle; reinforced
ness of confining reinforcement oriented either parallel or
concrete; side cover. perpendicular to the straight portion of hooked bars is also
evaluated to confirm whether current ACI 318 provisions
INTRODUCTION are appropriate.
Current design provisions for anchorage of hooked bars
in reinforced concrete (ACI Committee 318 2014; ACI EXPERIMENTAL INVESTIGATION
Committee 349 2006; AASHTO 2012) are based on several Tests of 166 simulated beam-column joint specimens
assumptions about their behavior: among others, hooked bars containing two hooked bars, included as part of a larger
with 90- and 180-degree bend angles are assumed to have research program (Sperry et al. 2015a,b; 2017a,b), were used
similar strengths, hooked bars with side covers of 2.5 in. to investigate the effect of bend angle, side cover, and orien-
(65  mm) or greater have similar strengths, and confining tation of confining reinforcement on anchorage strength.
reinforcement oriented parallel or perpendicular to the No. 5, 8, and 11 (No. 16, 25, and 36) hooked bars were
straight portion of a 90-degree hooked bar (“straight portion tested in normalweight concrete with compressive strengths
of the hooked bar” refers to the straight portion of the bar in ranging from 4300 to 16,500 psi (30 to 114 MPa). Nominal
the direction of the embedment or development length) is clear cover from the hooked bars to the outside of the column
assumed to be equally effective in providing confinement, (side cover) ranged from 2.5 to 3.5 in. (65 to 90 mm).
but only confining reinforcement oriented perpendicular to Bar stresses at anchorage failure ranged from 33,000 to
the straight portion of a 180-degree hooked bar is assumed 137,400 psi (228 to 947 MPa). The results of these tests are
to be effective. reported and used in conjunction with the results of previous
The design provisions are based on 38 tests by Marques studies (Marques and Jirsa 1975; Pinc et al. 1977; Hamad
and Jirsa (1975) and Pinc et al. (1977) of beam-column et al. 1993; Ramirez and Russell 2008; Lee and Park 2010)
joint specimens containing Grade 60 (420) No. 7, No. 9, or to determine the effect on hooked bar anchorage strength
No. 11 (No. 22, No. 29, or No. 36) bars with standard hooks of bend angle, concrete side cover, and confining reinforce-
(ACI Committee 318 2014), and concrete with compressive ment orientation. The details of the experimental investiga-
strengths ranging from 3600 to 5400 psi (25 to 37 MPa). tion are provided by Sperry et al. (2015a,b).
Marques and Jirsa (1975) observed that the thickness of the
concrete cover had a significant effect on the slip and stress Test specimens
at failure but indicated no advantage for covers greater than Figure 1 shows side and plan views of a typical specimen.
2.5 in. (65 mm). None of the test specimens in these earlier Specimens were designed to represent exterior beam-column
studies contained confining reinforcement perpendicular to ACI Structural Journal, V. 115, No. 1, January 2018.
the straight portion of the hooked bars. MS No. S-2017-075, doi: 10.14359/51700920, was received March 3, 2017, and
reviewed under Institute publication policies. Copyright © 2018, American Concrete
To validate the applicability of the earlier findings, tests Institute. All rights reserved, including the making of copies unless permission is
were performed in this study to evaluate the effects of obtained from the copyright proprietors. Pertinent discussion including author’s
closure, if any, will be published ten months from this journal’s date if the discussion
hook bend angle, concrete clear cover, and orientation of is received within four months of the paper’s print publication.

ACI Structural Journal/January 2018 247


Fig. 2—Specimen with confining reinforcement perpendic-
Fig. 1—Simulated beam-column specimens: (a) side view; ular to straight portion of hooked bar: (a) side view; and
(b) cross section without confining reinforcement; and (b) cross section (Sperry et al. 2017a).
(c) cross section of specimen with confining reinforce-
ment parallel to straight portion of hooked bar (Sperry et No. 3 (No. 10) hoops were used to confine No. 5 (No. 16)
al. 2017a). and No. 8 (No. 25) standard hooks and six No. 3 (No. 10)
hoops were used to confine No. 11 (No. 36) standard hooks.
joints and were cast without the beam. They were similar For case (3), the first hoop was centered 2db from the top
to the specimens used in the studies by Marques and Jirsa of the hooked bar (1.5db from the longitudinal axis of the
(1975) and Pinc et al. (1977). The specimens analyzed straight portion of the hooked bar). For case (2), in speci-
in this paper contained two hooked bars cast inside the mens with No. 5 or No. 8 hooked bars, the first hoop was
column longitudinal reinforcement. The out-to-out spacing spaced 3 in. (76.2 mm) from the center of the straight portion
of the hooked bars was the same for each bar size—8, 12, of the hooked bars; the second hoop was spaced 3 and 8 in.
and 16.5 in. (203, 305, and 419 mm) for specimens with (76.2 and 203.2 mm), respectively, from the first hoop. For
No. 5, No. 8, and No. 11 (No. 16, 25, and 36) hooked bars, case (2) in specimens with No. 11 hooked bars, the first and
respectively. Total column widths are given in Appendix A.* second hoops were spaced at 8 in. (203.2 mm) intervals from
In this paper, embedment length ℓeh refers to the distance the center of the straight portion of the hooked bars.
from the front of the column face to the back of the tail of To evaluate the effect of reinforcement orientation,
the hook. This is in contrast to the development length ℓdh, six specimens were tested with confining reinforcement
which refers to the minimum length of anchorage required oriented perpendicular to the straight portion of the hooked
in Section 25.4.3 of ACI 318-14 (ACI Committee 318 2014) bar, as shown in Fig. 2. Of the six, two contained two No. 3
to ensure that a bar can develop its yield strength. Column (No. 10) hoops, two contained four No. 3 (No. 10) hoops,
reinforcement was proportioned to resist the maximum and two contained five No. 3 (No. 10) hoops. The latter two
expected shear and moment assuming that both hooked bars cases meet the requirement to allow the use of ψr = 0.8.
reached their expected strength simultaneously. Specimen heights were chosen so that the support reac-
The specimens contained one of three quantities of tions from the test frame had minimal effect on the hook
confining reinforcement, in all but six cases oriented parallel region during testing, as shown in Fig. 3. The column height
to the straight portion of the hooked bar: 1) no confining was 52-3/4 in. (1340 mm) for the specimens with No. 5
reinforcement; 2) two No. 3 (No. 10) hoops spaced along (No. 16) or No. 8 (No. 25) hooked bars and 96 in. (2438 mm)
the length of the tail of the hook; or 3) No. 3 (No. 10) hoops for the specimens with No. 11 (No. 36) hooked bars. The
spaced at three bar diameters (3db) along the tail and the distances from the longitudinal axis of the straight portion of
bend of the hook, where db is the diameter of the hooked bar. the hooked bar to the upper compression member hcu were
No. 3 (No. 10) hoops spaced at 3db represents the amount 18.5, 18.5, and 48.5 in. (470, 470, and 1232 mm) for No. 5,
of confining reinforcement required in Section  25.4.3 No. 8, and No. 11 (No. 16, No. 25, and No. 36) hooked
of ACI 318-14 to allow the use of the confining reinforce- bars, respectively, and the distances from the center of the
ment modification factor ψr = 0.8 in the calculation of the hooked bar to the bearing member, simulating the compres-
development length of hooked bars. At this spacing, five sion region of the concrete beam framing into the column
hcl were 5.25, 10, and 19.2 in. (133, 254, and 488 mm) for
*
The Appendix is available at www.concrete.org/publications in PDF format, No. 5, No. 8, and No. 11 (No. 16, No. 25, and No. 36) hooked
appended to the online version of the published paper. It is also available in hard copy
from ACI headquarters for a fee equal to the cost of reproduction plus handling at the
bars, respectively.
time of the request.

248 ACI Structural Journal/January 2018


Test procedure
Specimens were tested using a self-reacting system config-
ured to simulate the boundary conditions of a beam-column
joint (Fig. 3). The test frame was a modified version of the
apparatus used by Marques and Jirsa (1975). The flange width
of the upper compression member and the bearing member
were 6-5/8 and 8-3/8 in. (168 and 213 mm), respectively.
For specimens with No. 5 and No. 8 (No. 16 and No. 25)
hooked bars, a constant column axial load of 30,000 lb
(133 kN) was applied to most of the specimens, corre-
sponding to a range in axial stress of 90 to 460 psi (0.6
to 3.2 MPa). For early tests, a constant force of 80,000 lb
(356 kN) was used, corresponding to an axial stress range
of 505 to 1930 psi (3.5 to 13.3 MPa) based on column size.
Specimens with No. 11 hooked bars had a constant axial
stress of 280 psi (1.9 MPa) applied. These axial stresses
were chosen based on the capacity of the axial load appli-
cation system. Marques and Jirsa (1975) found that changes
in axial stress up to 3000 psi (21 MPa) resulted in negligible
changes in the anchorage strength of the hooked bars; the
effect of varying axial stress was therefore not examined.
Fig. 3—Testing frame and forces applied to specimens
The load was applied monotonically to the hooked bars
during testing (Sperry et al. 2017b).
using hydraulic jacks to simulate tensile forces in the beam
Material properties reinforcement at the face of a beam-column joint. Specimens
Specimens were cast using non-air-entrained ready mixed were loaded to approximately 80% of the expected failure
concrete with nominal compressive strengths of 5000, 8000, load in, for most cases, four increments. Between incre-
12,000, and 15,000 psi (34, 55, 83, and 103 MPa). Measured ments, cracks were marked and photographs were taken.
strengths obtained using 6 x 12 in. (150 x 300 mm) cylin- Above 80% of the expected failure load, the specimens were
ders cured in the same manner as the test specimen and loaded continually to failure. Tests lasted 15 to 25 minutes.
tested on the same day as the anchorage tests ranged from Detailed descriptions of the test frame and testing procedure
4300 to 16,500 psi (30 to 114 MPa). The concrete contained are provided by Peckover and Darwin (2013).
Type I/II portland cement, crushed limestone, or granite
with a maximum size of 0.75 in. (19 mm), Kansas river RESULTS AND ANALYSIS
sand, and a high-range water-reducing admixture. Pea Results from 166 tests of beam-column joint specimens
gravel was incorporated as a portion of the aggregate in the with No. 5, No. 8, and No. 11 (No. 16, No. 25, and No. 36)
12,000 psi (83 MPa) concrete to improve the workability hooked bars were selected from the data presented by
of the mixture. Silica fume and Class C fly ash were used Sperry et al. (2015a,b, 2017b) to evaluate the effect of hook
as supplementary cementitious materials for the 15,000 psi bend angle, concrete side cover, and confining reinforce-
(103 MPa) concrete. Polycarboxylate-based high-range ment orientation on anchorage strength. These results were
water-reducing admixtures were used in the mixtures. combined with selected test results from Marques and Jirsa
Mixture proportions are presented in Appendix A and by (1975), Pinc et al. (1977), Hamad et al. (1993), Ramirez and
Sperry et al. (2015a,b, 2017a). Russell (2008), and Lee and Park (2010). The forces applied
Hooked bars were fabricated using ASTM A615 Grade 80 to the hooked bars at failure for the specimens included in
(550 MPa) and A1035 Grade 120 (830 MPa) reinforce- the analysis are presented in Appendix A. As described by
ment. For most specimens, the ancillary steel for column Sperry et al. (2017b), the average bar force at the peak load,
and confining reinforcement consisted of ASTM A615 equal to the maximum total force applied to a specimen
Grade 60 (420 MPa) reinforcing bars. ASTM A1035 Grade divided by the number of hooked bars under load, is treated
120 (830 MPa) bars were used as the column longitudinal as the failure load per hooked bar and used to calculate the
steel for specimens that had a greater flexural demand (based average bar stress at failure.
on the expected strength of the hooked bars) than could be To limit the effect of concrete compressive strength
satisfied using ASTM A615 Grade 60 (420 MPa) reinforcing on anchorage strength and simplify the comparisons, the
bars. Yield strength, nominal diameter, deformation spacing average bar forces at failure were normalized with respect
and height, gap width, and relative rib area for the deformed to a concrete compressive strength of 5000 psi (34.5 MPa)
steel bars used as hooked bars are presented in Appendix A by multiplying the average bar forces at failure T by
and in Sperry et al. (2017b). Appendix A also includes repre- (5000 fcm ) p 1 ( [34.5 fcm ] p 1). The result is reported as the
sentative stress-strain curves for the hooked bars. normalized average failure load TN. A value of p1 = 0.25 was
selected based on the observation by Sperry et al. (2015a,b;
2017a,b) that the power of 0.5, currently used in the provi-
sions of ACI 318-14 (ACI Committee 318 2014), greatly

ACI Structural Journal/January 2018 249


Fig. 4—Bar force at failure normalized to concrete compres- Fig. 5—Bar force at failure normalized to concrete compres-
sive strength of 5000 psi (34.5 MPa) versus embedment sive strength of 5000 psi (34.5 MPa) versus embedment length
length for No. 5, 7, 8, and 11 (No. 16, 22, 25, and 36) for No. 5 and 8 (No. 16 and 25) hooked bars confined by two
hooked bars without confining reinforcement. (Note: 1 kip = No. 3 (No. 10) hoops oriented parallel to straight portion of
4.448 kN; 1 in. = 25.4 mm.) hooked bar. (Note: 1 kip = 4.448 kN; 1 in. = 25.4 mm.)
overestimates the effect of concrete compressive strength are identified based on bar size. In this figure and those that
on anchorage strength, and that values of p1 in the range of follow, the order of the results in the legend coincides with
0.25 to 0.29 provide a realistic representation of the effect the order of the lines in the figure. For each bar size, the range
of concrete compressive strength on the anchorage strength of embedment lengths was similar for 90- and 180-degree
of hooked bars. This observation is further supported by test hooked bars. The embedment lengths ℓeh ranged from 6.3 to
results for straight bar lap splices (Darwin et al. 1995, 1996; 21.1 in. (160 to 536 mm), the bar stresses at failure ranged
Zuo and Darwin 1998, 2000; Joint ACI-ASCE Committee from 42,700 to 136,100 psi (294 to 938 MPa), and the
408 2003) and headed bars (Shao et al. 2016), which indicate normalized average bar forces at failure ranged from 19,300
that a value of p1 = 0.25 characterizes the effect of concrete to 114,400 lb (86 to 509 kN). The concrete compressive
compressive strength on both straight bar bond and headed strengths ranged from 2570 to 16,500 psi (17.7 to 114 MPa).
bar anchorage strength. As shown in Fig. 4, an increase in embedment length is
In the comparisons that follow, a regression analysis tech- associated with an increase in the normalized average bar
nique based on dummy variables (Draper and Smith 1981) force at failure. The results in Fig. 4 show no clear correla-
is used to identify trends in the data. Dummy variables anal- tion between anchorage strength and bend angle. For No. 5,
ysis is a least-squares regression analysis method that allows 7, and 11 (No. 16, 22, and 36) hooked bars, the trend line
differences in populations to be taken into account when corresponding to a 90-degree bend angle is higher than the
formulating relationships between principal variables. For trend line corresponding to a 180-degree bend angle. The
example, the effect of embedment length ℓeh on bar force opposite trend is observed for No. 8 (No. 25) hooked bars.
at failure T can be found for different bar sizes based on the The difference between intercepts of the trend lines corre-
assumption that the effect of changes in ℓeh on changes in T sponding to 90- and 180-degree bend angles is greater for
is the same for the bar sizes considered, but that the absolute the No. 11 (No. 36) bars than for the smaller bars. The results
value of T for a given ℓeh will differ for each bar size. were evaluated using Student’s t-test by comparing the inter-
cepts obtained by extending a line with the same slope as the
Effect of bend angle trend lines from each data point to the TN axis. That evalu-
Figure 4 shows the normalized average failure loads TN ation shows that the differences in anchorage strength for
as a function of embedment length for a subset of 58 beam- bars with bend angles of 90 and 180 degrees are not statis-
column specimens, of which 39 are from the current study. tically significant at the 95% confidence level (5% or lower
These specimens contained No. 5, No. 7, No. 8, and No. 11 risk of concluding that a difference exists when there is no
(No. 16, No. 22, No. 25, and No. 36) hooked bars without actual difference) regardless of bar size. In the cases shown
confining reinforcement in the joint region, with bend angles in Fig. 4, the variable p ranges between 0.13 and 0.80,
of 90 or 180 degrees. The results for the nine No. 7 (No. 22) where p less than 0.05 would indicate that the differences
hooked bar tests and 11 of the No. 11 (No. 36) hooked bar in anchorage strength based on bend angle are significant
tests were taken from studies by Marques and Jirsa (1975), (Wonnacott and Wonnacott 1985).
Pinc et al. (1977), Hamad et al. (1993), Ramirez and Russell The relationship between normalized failure load and
(2008), and Lee and Park (2010). The solid lines and data embedment length for 26 beam-column specimens (all from
points represent the results for 90-degree hooked bars, while the current study) containing No. 5 or No. 8 (No. 16 and
the broken lines and open data points represent the results 25) hooked bars with 90- or 180-degree hooks and with two
for 180-degree hooked bars. Both trend lines and data points No. 3 (No. 10) hoops in the joint region is shown in Fig. 5.

250 ACI Structural Journal/January 2018


Fig. 6—Bar force at failure normalized to concrete compres- Fig. 7—Bar force at failure normalized to concrete compres-
sive strength of 5000 psi (34.5 MPa) versus embedment sive strength of 5000 psi (34.5 MPa) versus embedment
length for No. 8 and 11 (No. 25 and 36) hooked bars with length for No. 5 (No. 16) hooked bars with different amounts
confining reinforcement conforming to Section 25.4.3.2 of No. 3 (No. 10) bar confining reinforcement oriented
of ACI 318-14 and oriented parallel to straight portion of parallel to straight portion of hooked bar and side cover
hooked bar. (Note: 1 kip = 4.448 kN; 1 in. = 25.4 mm.) (sc). (Note: 1 kip = 4.448 kN; 1 in. = 25.4 mm.)

The two hoops were oriented parallel to the straight portion to 594 kN), the bar stresses at failure ranged from 69,400 to
of the hooked bars. Two hoops are insufficient to satisfy ACI 113,900 psi (478 to 785 MPa), and the concrete compressive
Code (ACI 318-14) requirements for the use of a develop- strengths ranged from 5420 to 15,800 psi (37.4 to 109 MPa).
ment length modification factor ψr = 0.8 for hooked bars, and For both the No. 8 and No. 11 (No. 25 and No. 36) hooked
hoops oriented parallel to the straight portion of the hooked bars, the anchorage strength of the hooked bars with a bend
bar, regardless of number or spacing, are not considered by angle of 180 degrees was slightly lower than the strength of
the Code to increase the anchorage strength of bars with the hooked bars with a 90-degree bend angle, although the
180-degree hooks. Contrary to the Code, however, Sperry results of Student’s t-test show that the differences are not
et al. (2015a,b; 2017a,b) have shown that hoops placed statistically significant (p = 0.54 and 0.50, respectively).
parallel to the straight portion of hooked bars provide similar Because differences between the anchorage strengths
increases in anchorage strength for both 90- and 180-degree of hooked bars with 90- and 180-degree bend angles were
hooked bars. The embedment lengths ℓeh ranged from 5.6 to found to be small and not statistically significant, hooked
17.3 in. (142 to 439 mm), the normalized average bar forces bars with either bend angle, and with all other parameters
at failure TN ranged from 20,000 to 87,500 lb (89 to 389 kN), the same, should be treated as having the same anchorage
the bar stresses at failure ranged from 68,000 to 137,400 psi strength, as reflected in the design provisions of ACI 318-14.
(469 to 947 MPa), and the concrete compressive strengths Further, confining reinforcement parallel to the straight
ranged from 4300 to 15,800 psi (30 to 109 MPa). portion of hooks is shown to provide the same contribution
The trend lines for anchorage strength nearly coincide for to the anchorage strength for both 90- and 180-degree stan-
the 90- and 180-degree No. 5 (No. 16) hooked bars, while the dard hooks.
180-degree No. 8 (No. 25) hooked bars had a lower strength
than the 90-degree No. 8 (No. 25) hooked bars. The results Effect of side cover
of Student’s t-test show that the differences in anchorage Based on the observations that bend angle has no measur-
strength for No. 5 and 8 (No. 16 and 25) bars with 90- or able effect on anchorage strength, the comparisons in this
180-degree hooks are not statistically significant at the 95% section include specimens with both 90- and 180-degree
confidence level, with p = 0.81 and 0.12, respectively. bend angles. The relationship between normalized anchorage
Figure 6 compares the anchorage strengths of No. 8 strength TN and embedment length ℓeh for 39 beam-column
and No. 11 (No. 25 and 36) bars with 90- and 180-degree joint specimens containing No. 5 (No. 16) hooked bars is
hooks confined by No. 3 (No. 10) hoops oriented parallel shown in Fig. 7. The specimens in Fig. 7 were tested as part
to the straight portion of the hooked bar and spaced at 3db, of this study and had nominal side covers of 2.5 in. (65 mm)
which satisfies the requirements for use of the development (solid lines) or 3.5 in. (90 mm) (broken lines). Three different
length modification factor ψr = 0.8 for 90-degree hooks in quantities of confining reinforcement (parallel to the straight
ACI 318-14, Section 25.4.3.2. The data in Fig. 6 represent portion of the hooked bar) were investigated: no confining
18 specimens, all tested as part of this study, containing reinforcement; two No. 3 (No. 10) hoops within the joint
No. 8 or No. 11 (No. 25 or No. 36) hooked bars with 90- and region; and No. 3 (No. 10) hoops spaced at 3db (satisfying
180-degree bend angles. The embedment lengths ℓeh ranged the requirements for ψr = 0.8). The embedment lengths ℓeh
from 9.4 to 20.4 in. (239 to 518 mm), the normalized average ranged from 3.75 in. to 10.5 in. (95 to 267 mm). The normal-
bar forces at failure TN ranged from 51,700 to 133,600 lb (230 ized failure load TN ranged from 14,000 to 41,500 lb (62 to

ACI Structural Journal/January 2018 251


Fig. 8—Bar force at failure normalized to concrete compres- Fig. 9—Bar force at failure normalized to concrete compres-
sive strength of 5000 psi (34.5 MPa) versus embedment sive strength of 5000 psi (34.5 MPa) versus embedment length
length for No. 8 (No. 25) hooked bars with different amounts for No. 11 (No. 36) hooked bars with different amounts of
of No. 3 (No. 10) bar confining reinforcement oriented No. 3 (No. 10) bar confining reinforcement oriented parallel
parallel to straight portion of hooked bar and side cover to straight portion of hooked bar and side cover (sc). (Note:
(sc). (Note: 1 kip = 4.448 kN; 1 in. = 25.4 mm.) 1 kip = 4.448 kN; 1 in. = 25.4 mm.)
185 kN), the bar stresses at failure ranged from 60,300 to specimens with two No. 3 (No. 10) hoops and No. 3 (No.
136,100 psi (416 to 938 MPa), and the concrete compressive 10) hoops spaced at 3db are not statistically significant at the
strengths ranged from 5190 to 15,800 psi (36 to 109 MPa). The 95% confidence level, with p equal to 0.32 and 0.47, respec-
trend lines in Fig. 7 show that anchorage strength increased tively. For specimens without confining reinforcement, the
with increasing embedment length and amount of confining difference in anchorage strength between hooked bars with
reinforcement. Regardless of the amount of confining rein- 2.5 or 3.5 in. (65 or 90 mm) side cover is statistically signif-
forcement, the results shown in Fig. 7 indicate that there was icant (p = 0.03).
a decrease in strength as the side cover increased from 2.5 The relationship between normalized bar force at failure
to 3.5 in. (65 to 90 mm), although the results from Student’s TN and embedment length ℓeh is shown in Fig. 9 for 49 beam-
t-test show that this decrease was not statistically significant column joint specimens with No. 11 (No. 36) hooked bars
at the 95% confidence level, with p = 0.72, 0.08, and 0.30 with nominal side covers of 2.5 or 3.5 in. (65 or 90 mm)
for specimens without confining reinforcement, specimens tested as part of this study. The embedment lengths ℓeh
with two No. 3 (No. 10) hoops, and specimens with No. 3 ranged from 9.5 to 26.0 in. (241 to 660 mm), the concrete
(No. 10) hoops spaced at 3db, respectively. compressive strengths ranged from 4910 to 16,180 psi (34 to
The relationship between normalized failure load TN and 112 MPa), the bar stresses at failure ranged from 33,000 to
embedment length ℓeh for beam-column joint specimens with 136,700 psi (228 to 943 MPa), and the normalized average
No. 8 (No. 25) hooked bars with nominal side covers of 2.5 bar forces at failure TN ranged from 39,800 to 174,400 lb
or 3.5 in. (65 or 90 mm) is shown in Fig. 8. The data represent (177 to 776 kN). Similar to the test results for the smaller
78 specimens tested as part of this study. The embedment hooked bars, anchorage strength increased with embed-
lengths ℓeh ranged from 6.1 to 18.7 in. (155 to 475 mm), the ment length and the amount of confining reinforcement. For
concrete compressive strengths ranged from 4300 to 16,500 specimens without confining reinforcement and specimens
psi (30 to 114 MPa), the bar stresses at failure ranged from with two No. 3 (No. 10) hoops, the change in the normal-
44,430 to 120,700 psi (306 to 832 MPa), and the normal- ized failure load was negligible as side cover increased from
ized failure load TN ranged from 28,200 to 93,600 lb (125 to 2.5 to 3.5 in. (65 to 90 mm). For specimens with six No. 3
416 kN). The trend lines in Fig. 8 show that the normalized (No. 10) hoops spaced at 3db, the trend lines show that there
failure load increased with increasing embedment length and was a small decrease in normalized failure load as side cover
amount of confining reinforcement. For specimens without increased from 2.5 to 3.5 in. (65 to 90 mm). Because there
confining reinforcement, increasing the side cover from was only one specimen with 3.5 in. (90 mm) side cover, the
2.5 to 3.5 in. (65 to 90 mm) led to increases in anchorage most meaningful conclusion that can be drawn from the data
strength. Specimens with confining reinforcement included is that the normalized failure load was comparable for the
configurations with two No. 3 (No. 10) hoops and with five two configurations. The results of Student’s t-test indicate
No. 3 (No. 10) hoops spaced at 3db in the joint region. For that the differences in anchorage strength associated with
both of these configurations, specimens with 3.5  in. (90 changes in side cover for specimens without confining rein-
mm) side cover had normalized failure loads slightly lower forcement and specimens with two No. 3 (No. 10) hoops
than those of specimens with 2.5 in. (65 mm) side cover. are not statistically significant at the 95% confidence level
The results of Student’s t-test show that the differences in (p = 0.56 and 0.82, respectively). Student’s t-test cannot
anchorage strength associated with changes in cover for be performed for the specimens with No. 3 (No. 10) hoops

252 ACI Structural Journal/January 2018


Fig. 10—Plan view of hooked bars with confining reinforcement oriented perpendicular to bar being developed satisfying
maximum spacing requirement in ACI 318-14, Section 25.4.3.2: (a) four No. 3 (No. 10) hoops; and (b) five No. 3 (No. 10)
hoops. (Note: 1 in. = 25.4 mm.)
spaced at 3db because there was only one specimen with contained hooked bars with a 180-degree bend angle. For
3.5 in. (90 mm) side cover. both sets of six, one specimen was cast without confining
Of all the Student’s t-tests performed, in only one instance, reinforcement, one contained two hoops oriented parallel to
No. 8 (No. 25) hooked bars without confining reinforcement the bars being developed (parallel hoops), one contained two
was the value of p less than 0.05, indicative of a statistically hoops oriented perpendicular to the bars being developed
significant difference between the anchorage strength of (perpendicular hoops), one contained parallel hoops spaced
hooked bars with 2.5 in. (65 mm) side cover and hooked bars at 3db, and two contained perpendicular hoops spaced at less
with 3.5 in. (90 mm) side cover. Overall, the results indicate than 3db.
that anchorage strength was not affected by differences in In the specimens with parallel hoops placed along the tail
side cover in the range of 2.5 to 3.5 in. (65 to 90 mm). of the hook, a minimum of five hoops were needed to meet
the 3db spacing requirement. In the specimens with perpen-
Effect of orientation of confining reinforcement dicular hoops, the specimen dimensions and the depth of the
To take advantage of the modification factor ψr = 0.8 for joint were such that only four hoops were needed to meet
development length of hooked bars with a 90-degree bend the 3db spacing requirement specified in the code provisions
angle, ACI 318-14, Section 25.4.3.2, requires confining (ACI 318-14, Section 25.4.3.2). To evaluate the effect of
reinforcement spaced at ≤3db and placed perpendicular or hoop orientation on anchorage strength without favoring the
parallel to the straight portion of the bar being developed. parallel orientation, two different reinforcement configura-
For hooked bars with a 180-degree bend angle, ψr = 0.8 can tions were used in specimens with perpendicular hoops satis-
only be applied for confining reinforcement oriented perpen- fying the 3db maximum spacing requirement—one with the
dicular to the straight portion of the bar. The equivalence of minimum of four hoops needed to achieve the 3db spacing
the contribution of confining reinforcement oriented parallel and one with five hoops to match the area of confining rein-
to the straight portion of a hooked bar to the anchorage forcement used in the specimens with parallel hoops spaced
strength of bars with 90- and 180-degree hooks has already at 3db. The two configurations are shown in Fig. 10. For
been demonstrated in this paper. Prior to this study, no tests specimens with a 180-degree bend angle, parallel hoops
had been performed to evaluate the contribution to anchorage were placed throughout the region defined by the bend and
strength of confining reinforcement oriented perpendicular tail of a 90-degree hooked bar, as shown in Fig. 1.
to the straight portion of a hooked bar. Because all of the specimens in this batch had different
To address this question, a test series was planned with embedment lengths ℓeh and slightly different concrete
12 similar beam-column joint specimens cast with the same compressive strengths fcm, the average bar forces at failure
batch of concrete. This set of specimens contained No. 8 T were normalized with respect to an embedment length
(No. 25) hooked bars with bend angles of 90 or 180 degrees. of 10 in. (254 mm) and a concrete compressive strength of
The specimens had embedment lengths ℓeh with nominal 12,000 psi (83 MPa) by multiplying the average bar forces at
values of 10, 11, and 12.5 in. (254, 279, and 318 mm). The
failure T by (10  eh )(12,000 f cm ) ( 254  eh )(83 f cm ) 
0.25 0.25

corresponding column cross-sectional dimensions were


to obtain the normalized average failure load TN.
17 x 12 in., 17 x 13 in., and 17 x 14.5 in. (432 x 305 mm,
The values of TN for the specimens with two No. 3
432 x 330 mm, and 432 x 368 mm). Compressive strengths
(No. 10) hoops and the specimens with No. 3 (No. 10) hoops
on the day of the tests were similar, ranging from 11,800 to
spaced at ≤3db are shown in the bar graph in Fig. 11. The
12,000 psi (81 to 83 MPa). Average measured embedment
first set of four bars in the graph shows TN for the 90- and
lengths ranged from 9.4 to 12.8 in. (234 to 325 mm), bar
180-degree hooked bars confined by two hoops oriented
stresses at failure ranged from 66,600 to 95,200 psi (459 to
parallel or perpendicular to the straight portion of the bar.
656 MPa), and average failure loads ranged from 60,200
As shown for these four specimens, the value of TN for the
to 75,200 lb (268 to 335 kN). Of the 12 specimens, six
specimen containing 90-degree hooked bars confined by
contained hooked bars with a 90-degree bend angle and six
parallel hoops was 1.27 times the value of TN for the spec-

ACI Structural Journal/January 2018 253


Fig. 11—Bar force at failure normalized to embedment Fig. 12—Ratio of anchorage strength of No. 8 (No. 25)
length of 10 in. (254 mm) and concrete compressive strength hooked bars with confining reinforcement oriented parallel
of 12,000 psi (83 MPa) for specimens containing No. 8 to straight portion of bar to anchorage strength of No. 8
(No.  25) hooked bars with No. 3 (No. 10) confining rein- (No. 25) hooked bars with confining reinforcement oriented
forcement oriented parallel or perpendicular to straight perpendicular to straight portion of bar. Strengths normal-
portion of hooked bar and 90- or 180-degree bend angles. ized to embedment length of 10 in. (254 mm) and concrete
(Note: 1 kip = 4.448 kN.) compressive strength of 12,000 psi (83 MPa).
imen containing 90-degree hooked bars confined by perpen- with, respectively, four and five perpendicular hoops. In this
dicular hoops. For the specimens containing 180-degree limited test series, increasing from four to five perpendicular
hooked bars, TN for the specimens with the two orientations hoops had little effect on the strength of hooked bars with
of confining reinforcement had similar strengths—TN for the bend angles of either 90 or 180 degrees.
specimen with perpendicular hoops was 1.01 times TN for Figure 12 shows the ratio of the anchorage strength of the
the specimen with parallel hoops. hooked bars confined by parallel hoops to the anchorage
The second and third sets of four bars in Fig. 11 show the strength of hooked bars confined by perpendicular hoops.
results for specimens with hoops spaced ≤3db. In each of The figure indicates that for 90-degree hooked bars, all of
these sets, two specimens were cast with hoops oriented in which had ratios greater than unity, parallel hoops provided
the parallel direction, one with 90-degree hooked bars and a greater increase in anchorage strength than perpendicular
the other with 180-degree hooked bars, both having five hoops. In contrast, for 180-degree hooked bars, all had ratios
hoops spaced ≤3db. Because there were no specimens with below unity, indicating that perpendicular hoops provided
four hoops with a parallel orientation, specimens with five the greater increase in strength. The greater contribution
hoops are shown in the first and third columns of the second of parallel hoops for 90-degree hooked bars may result
set for reference. The trends for the specimens with hoops because the hoops serve as anchor reinforcement that resists
spaced ≤3db are similar to those observed for specimens with a breakout failure. In this configuration, hoop orientation is
two hoops. The value of TN for the specimen with 90-degree optimal for carrying a direct tensile force, which helps keep
hooked bars confined by parallel hoops is higher than TN for the joint region adjacent to the 90-degree hooked bar intact.
the specimens containing 90-degree hooked bars confined In contrast, perpendicular hoops do not develop a direct
by perpendicular hoops, although the difference is smaller tensile force to counteract the tensile force in the hooked
than observed for the specimens with two hoops. TN for the bars, causing the hoops to be less efficient than hoops with
90-degree hooked bar specimen with five parallel hoops a parallel orientation in preventing a breakout failure of the
was, respectively, 1.21 and 1.16 times TN for the 90-degree joint region. Confining reinforcement oriented perpendic-
hooked bar specimens with four and five perpendicular ular to the straight portion of the hooked bar, however, may
hoops. For the 180-degree hooked bar specimens, the spec- be more efficient in limiting splitting of the concrete caused
imen with four perpendicular hoops had the highest strength, by slip of the hooked bars—splitting that may be greater for
while the specimen with five perpendicular hoops had the 180-degree hooked bars than for 90-degree hooked bars.
lowest. TN for the 180-degree hooked bar specimen with five Greater slip was observed for 180-degree hooked bars by
parallel hoops was, respectively, 0.97 and 1.05 times the Marques and Jirsa (1975) and Hamad et al. (1993). Splitting
failure loads of the companion specimens with four and five due to slip is also key in developing straight bar reinforce-
perpendicular hoops. TN for the 180-degree hooked bar spec- ment, where the resistance to the wedging action of the bar
imen with five parallel hoops was equal to 0.95 of TN for the due to slip is a function of the amount of confining rein-
90-degree hooked bar specimen with parallel hoops and 1.15 forcement oriented perpendicular to the bar and the concrete
and 1.11 times TN for the 90-degree hooked bar specimens compressive strength. This suggests that the confinement

254 ACI Structural Journal/January 2018


provided by perpendicular hoops may be similar to that of specimens with perpendicular hoops, the average anchorage
the confinement provided by reinforcement perpendicular strength ratio was approximately 0.85, which shows that
to straight bars (Darwin et al. 1995, 1996; Zuo and Darwin placing confining reinforcement perpendicular to the
1998, 2000, Joint ACI-ASCE Committee 408 2003). straight portion of the hooked bar resulted in lower increases
Figure 13 shows the ratio of the anchorage strength of in anchorage strength for 90-degree hooked bars than for
90-degree hooked bars to that of 180-degree hooked bars for 180-degree hooked bars.
specimens with both hoop orientations. The ratio for speci- The observed failure modes (Fig. 14(a) and (b)) support
mens with parallel hoops ranged from 1.01 to 1.06, while the the theory that confining reinforcement oriented parallel
ratio for specimens with perpendicular hoops ranged from to the straight portion of a hooked bar is more effective in
0.80 to 0.89. For specimens with parallel hoops, anchorage keeping the concrete in the joint region together by acting as
strength ratios were very close to 1.0, which indicates that anchor reinforcement, limiting the potential for a concrete
regardless of the number of hoops in the specimens, placing breakout failure at the front of the column. As observed by
the confining reinforcement parallel to the straight portion Searle et al. (2014) and Sperry et al. (2015a,b; 2017a), the
of the hooked bar resulted in a similar increase in strength tensile force in the hooked bars pulls out a triangular-shaped
for hooked bars with 90- and 180-degree bend angles. For block of concrete (with its base at the front of the column),
as shown in Fig. 14(a), with the tensile force in the hoops
directly opposing that breakout force and pinning the trian-
gular-shaped block to the compression zone at the back of
the column. When reinforcement oriented perpendicular to
the straight portion of the bar is used to confine hooked bars
with 90-degree bend angles, it does help to keep the concrete
in the joint together, although it is less effective because it
no longer acts as anchor reinforcement. Failure occurs as a
block of confined concrete is pulled toward the front of the
column with the hoops, as shown in Fig. 14(b).
Based on these results, it is concluded that the anchorage
strength of 180-degree hooked bars with either orientation
of confining reinforcement is similar to that of 90-degree
hooked bars with confining reinforcement oriented parallel to
the straight portion of the hooked bar. Confining reinforce-
ment oriented perpendicular to the straight portion of the bar
is less effective for 90-degree hooked bars. Considering the
limited amount of test data in this study to address the effect
Fig. 13—Ratio of anchorage strength of No. 8 (No. 25) hooked
on anchorage strength of confining reinforcement perpen-
bars with 90-degree bend angle to anchorage strength of
dicular to the straight portion of a hooked bar, more research
No. 8 (No. 25) hooked bars with 180-degree bend angle for
on the effect of confining reinforcement with this orientation
hooks confined by No. 3 (No. 10) hoops oriented parallel or
is needed.
perpendicular to straight portion of bar. Strengths normal-
ized to embedment length of 10 in. (254 mm) and concrete
compressive strength of 12,000 psi (83 MPa).

Fig. 14—Failure modes of specimens with 90-degree hooked bars with different orientations of confining reinforcement: (a)
hoops parallel to straight portion of hooked bars pinning triangular-shaped block to compression zone; and (b) hoops perpen-
dicular to straight portion of hooked bar with concrete block being pulled toward front of column.

ACI Structural Journal/January 2018 255


SUMMARY AND CONCLUSIONS ACI member Nathaniel Searle is a Structural Design Engineer with S. A.
Miro, Inc. in Denver, CO. He received his BS and MS degrees from the
Test results for 166 simulated exterior beam-column joints University of Kansas.
with two hooked bars were used to investigate the effects
of bend angle, concrete side cover, and confining reinforce- ACI member Michael DeRubeis is a Structural Design Engineer with S. A.
Miro, Inc., in Denver, CO. He received his BS from the University of Mich-
ment orientation on the anchorage strength of hooked bars. igan, Ann Arbor, MI, and his MS from the University of Kansas.
The simulated beam-column joint specimens contained
No. 5, No. 8, and No. 11 (No. 16, No. 25, and No. 36) ACI member Ali Ajaam is a Faculty Member at the University of Babylon,
Babylon, Iraq. He received his BSc and MSc from the University of Babylon
hooked bars with 90- or 180-degree bend angles. The clear and his PhD from the University of Kansas.
concrete side cover ranged from 2.5 to 3.5 in. (65 to 90 mm).
The specimens were cast using normalweight concrete with ACKNOWLEDGMENTS
compressive strengths ranging from 4300 to 16,500 psi (30 Support for the study was provided by the Electric Power Research Insti-
tute, Concrete Reinforcing Steel Institute Education and Research Foun-
to 114 MPa). Bar stresses at failure ranged from 33,000 dation, University of Kansas Transportation Research Institute, Charles
to 137,400 psi (228 to 947 MPa). A set of specimens was Pankow Foundation, Commercial Metals Company, Gerdau Corporation,
cast with confining reinforcement oriented either parallel or Nucor Corporation, and MMFX Technologies Corporation. Additional
materials were supplied by Dayton Superior, Midwest Concrete Materials,
perpendicular to the straight portion of the hooked bar in and W. R. Grace Construction. Thanks are due to K. Barry and M. Ruis,
the joint region, with all other parameters held constant to who provided project oversight for the Advanced Nuclear Technology
study the effect of confining reinforcement orientation on Program of the Electric Power Research Institute, and to N. Anderson,
C. Kopczynski, M. Mota, J. Munshi, and C. Paulson, who served as
anchorage strength. Data from prior studies were included in industry advisors.
the analysis where applicable.
The following conclusions are based on the data and REFERENCES
analysis presented herein: AASHTO, 2012, AASHTO LRFD Bridge Design Specifications, sixth
1. Hooked bars with 90- and 180-degree bend angles have edition, American Association of State Highway and Transportation Offi-
cials, Washington, DC, 1672 pp.
similar anchorage strengths and can be used interchange- ACI Committee 318, 2014, “Building Code Requirements for Struc-
ably. This includes hooked bars with a 180-degree bend tural Concrete (ACI 318-14) and Commentary (ACI 318R-14),” American
angle confined by parallel reinforcement following Section Concrete Institute, Farmington Hills, MI, 519 pp.
ACI Committee 349, 2006, “Code Requirements for Nuclear Safety
25.4.3.2 of ACI 318-14 to allow use of the 0.8 modification Related Concrete Structures (ACI 349-06),” American Concrete Institute,
factor for calculating the development length of hooked bars. Farmington Hills, MI, 157 pp.
2. For hooked bars with a 90-degree bend angle, confining ASTM A1035/A1035M, 2014, “Standard Specification for Deformed
and Plain Low-Carbon, Chromium, Steel Bars for Concrete Reinforce-
reinforcement placed perpendicular to the straight portion of ment,” ASTM International, West Conshohocken, PA, 7 pp.
the bars results in a lower anchorage strength than confining ASTM A615/A615M, 2015, “Standard Specification for Deformed and
reinforcement with a similar spacing placed parallel to the Plain Carbon-Steel Bars for Concrete Reinforcement,” ASTM Interna-
tional, West Conshohocken, PA, 8 pp.
straight portion of the bars. Darwin, D.; Tholen, M. L.; Idun, E. K.; and Zuo, J., 1995, “Splice
3. Increasing concrete side cover from 2.5 to 3.5 in. (65 Strength of High Relative Rib Area Reinforcing Bars,” SL Report 95-3,
to 90 mm) does not increase the anchorage strength of University of Kansas Center for Research, Lawrence, KS, May, 58 pp.
Darwin, D.; Zuo, J.; Tholen, M. L.; and Idun, E. K., 1996, “Development
hooked bars. Length Criteria for Conventional and High Relative Rib Area Reinforcing
Bars,” ACI Structural Journal, V. 93, No. 3, May-June, pp. 347-359.
AUTHOR BIOS Draper, N. R., and Smith, H., 1981, Applied Regression Analysis, second
ACI member Jayne Sperry is a Graduate Engineer with Walter P Moore edition, Wiley, New York, 709 pp.
in Orlando, FL. She received her BS, MS, and PhD from the University of Hamad, B. S.; Jirsa, J. O.; and D’Abreu de Paulo, N. I., 1993, “Effect of
Kansas, Lawrence, KS. Epoxy Coating on Bond Anchorage of Reinforcing in Concrete Structures,”
ACI Structural Journal, V. 90, No. 1, Jan.-Feb., pp. 77-88.
ACI Honorary Member David Darwin is the Deane E. Ackers Distin- Joint ACI-ASCE Committee 408, 2003, “Bond and Development of
guished Professor and Chair of the Department of Civil, Environmental, Straight Reinforcing Bars in Tension (ACI 408R-03),” American Concrete
and Architectural Engineering at the University of Kansas. Institute, Farmington Hills, MI, 8 pp.
Lee, J., and Park, H., 2010, “Bending—Applicability Study of Ultra-Bar
ACI member Matthew O’Reilly is an Assistant Professor of civil, environ- (SD 600) and Ultra-Bar for Rebar Stirrups and Ties (SD 500 and 600) for
mental, and architectural engineering at the University of Kansas. Compression Rebar,” Korea Concrete Institute, Aug., 504 pp. (translated
from Korean)
Andrés Lepage, FACI, is an Associate Professor and Director of Labora- Marques, J. L., and Jirsa, J. O., 1975, “A Study of Hooked Bar Anchor-
tories in the Department of Civil, Environmental, and Architectural Engi- ages in Beam-Column Joints,” ACI Journal Proceedings, V. 72, No. 5, May,
neering at the University of Kansas. pp. 198-209.
Peckover, J., and Darwin, D., 2013, “Anchorage of High-Strength Rein-
ACI member Rémy D. Lequesne is an Assistant Professor of civil, environ- forcing Bars with Standard Hooks: Initial Tests” SL Report 13-1, University
mental, and architectural engineering at the University of Kansas. of Kansas Center for Research, Lawrence, KS, 47 pp.
Pinc, R.; Watkins, M.; and Jirsa, J. O., 1977, “The Strength of the Hooked
Adolfo Matamoros, FACI, is the Peter T. Flawn Distinguished Professor Bar Anchorages in Beam-Column Joints,” CESRL Report No. 77-3, Depart-
of Civil and Environmental Engineering at the University of Texas at San ment of Civil Engineering-Structures Research Laboratory, University of
Antonio, San Antonio, TX. Texas at Austin, Austin, TX, 67 pp.
Ramirez, J. A., and Russell, B. W., 2008, Transfer, Development, and
Lisa R. Feldman, FACI, is an Associate Professor of civil and geological Splice Length for Strand/reinforcement in High-strength Concrete, Trans-
engineering at the University of Saskatchewan, Saskatoon, SK, Canada, portation Research Board, National Research Council, Washington, DC,
and Director of the Saskatchewan Centre for Masonry Design. pp. 99-120.
Searle, N.; DeRubeis, M.; Darwin, D.; Matamoros, A.; O’Reilly, M.; and
ACI member Samir Yasso is a Faculty Member at the University of Mosul, Feldman, L., 2014, “Anchorage of High-Strength Reinforcing Bars with
Mosul, Iraq. He received his BSc and MSc from the University of Mosul and Standard Hooks - Initial Tests,” SM Report No. 108, University of Kansas
his PhD from the University of Kansas. Center for Research, Inc., Lawrence, KS, Feb., 120 pp.

256 ACI Structural Journal/January 2018


Shao, Y.; Darwin, D.; O’Reilly, M.; Lequesne, R. D.; Ghimire, K.; and Strength Hooked Bars—Part 2: Data Analysis,” ACI Structural Journal,
Hano, M., 2016, “Anchorage of Conventional and High-Strength Headed V. 114, No. 1, Jan.-Feb., pp. 267-276. doi: 10.14359/51689457
Reinforcing Bars,” SM Report No. 117, University of Kansas Center for Sperry, J.; Yasso, S.; Searle, N.; DeRubeis, M.; Darwin, D.; O’Reilly,
Research, Lawrence, KS, Aug., 234 pp. M.; Matamoros, A.; Feldman, L.; Lepage, A.; and Lequesne, R., 2017b,
Sperry, J.; Al-Yasso, S.; Searle, N.; DeRubeis, M.; Darwin, D.; O’Reilly, M.; “Conventional and High-Strength Hooked Bars—Part 1: Anchorage
Matamoros, A.; Feldman, L.; Lepage, A.; Lequesne, R.; and Ajaam, A., 2015a, Tests,” ACI Structural Journal, V. 114, No. 1, Jan.-Feb., pp. 255-266. doi:
“Anchorage of High-Strength Reinforcing Bars with Standard Hooks,” SM 10.14359/51689457
Report No. 111, University of Kansas Center for Research, Inc., Lawrence, Wonnacott, R., and Wonnacott, T., 1985, Introductory Statistics, Wiley,
KS, June 2015, 266 pp. New York, 649 pp.
Sperry, J.; Darwin, D.; O’Reilly, M.; and Lequesne, R., 2015b, Zuo, J., and Darwin, D., 1998, “Bond Strength of High Relative Rib
“Anchorage Strength of Conventional and High-Strength Hooked Bars in Area Reinforcing Bars,” SM Report No. 46, University of Kansas Center
Concrete,” SM Report No. 115, University of Kansas Center for Research, for Research, Lawrence, KS, Jan., 350 pp.
Lawrence, KS, Dec., 266 pp. Zuo, J., and Darwin, D., 2000, “Splice Strength of Conventional and
Sperry, J.; Darwin, D.; O’Reilly, M.; Matamoros, A.; Feldman, L.; High Relative Rib Area Bars in Normal and High-Strength Concrete,” ACI
Lepage, A.; Lequesne, R.; and Yasso, S., 2017a, “Conventional and High- Structural Journal, V. 97, No. 4, July-Aug., pp. 630-641.

ACI Structural Journal/January 2018 257


Regular Price: $69.50
Member Price: $39.00

Only available in PDF version

44
• Design guidelines and specifications;
• Material properties for design;
• Behavior and design of beams and
columns;
• Behavior and design of slabs and
other structures;
• Behavior and design of foundations and
papers are organized underground components; and finally,

6
• Applications in structure and under-
into themes: ground construction projects.

Held at Polytechnique Montreal, Canada, on July 24 and 25, 2014, the workshop
demonstrated the state-of-the-art progress attained in terms of specifications
and applications of fiber-reinforced concrete.

| +1.248.848.3700 | www.concrete.org |
ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 115-S21

Thermomechanical Relaxation of Reinforced Concrete


Beams Strengthened with Carbon Fiber-Reinforced
Polymer
by Yail J. Kim and Abdulaziz Alqurashi

This paper presents the behavior of reinforced concrete beams alent sections can be used to represent the effect of insulation
strengthened with carbon fiber reinforced polymer (CFRP) sheets layers. The CFRP-concrete interface is degraded by thermal
subjected to relaxation induced by thermal and mechanical load- loading primarily due to a change in the tangled cross-
ings that are applied simultaneously (thermomechanical loading). linking structure of polymeric adhesives.3 The interfacial
The range of elevated temperatures varies from 75 to 150°C (167 to
behavior is thus concurrently controlled by temperature and
302°F) and the mechanical load represents a typical service condi-
exposure time. Gamage et al.4 examined the characteristics
tion (60% of the strengthened beam’s capacity at 25°C [77°F]),
both of which are sustained for 30- to 75-minute periods. The heat of CFRP sheets bonded to a concrete substrate at elevated
conduction of the beams is measured by an infrared imaging tech- temperatures. The interface was insulated and subjected
nique to understand their thermal responses. Upon completion to temperatures from 600 to 1200°C (1112 to 2192°F) for
of the thermomechanical relaxation test, all beams are loaded to up to 3 hours in an oven. The interfacial capacity was not
failure for residual capacity investigations. The progression of controlled by CFRP-bond length, but was affected by the
thermal loading across the CFRP-concrete interface exhibits a thermal properties of insulation. Leone et al.5 evaluated
so-called latent heat effect that takes place with thermal conduction the bond of CFRP-concrete interface at elevated tempera-
along the interface. During residual testing, the degraded interface tures varying from 20 to 80°C (68 to 176°F). The interfacial
reveals partial debonding that accompanies the retraction of CFRP capacity was remarkably reduced when the applied tempera-
strains. Thermomechanical relaxation influences the residual
ture exceeded the adhesive’s glass transition temperature of
capacity, load displacement, and interfacial cracking patterns of
55°C (131°F), which was associated with a state change from
the strengthened beams. According to analytical modeling, the
elastic modulus of CFRP and the magnitude of a mechanical load solid to amorphous. The failure mode of the interface was
affect the degree of thermomechanical relaxation. A design factor is also influenced by temperatures: concrete cohesion failure at
proposed based on multiple regression to quantify the implications 50°C (92°F) and adhesive adhesion failure at 80°C (176°F).
of thermomechanical relaxation for CFRP-strengthened beams. It was recommended that thermal loading be considered in
CFRP-strengthening design.
Keywords: carbon fiber-reinforced polymer (CFRP); relaxation; retrofit; Structural materials experience stress relaxation in
strengthening; thermal.
prescribed boundary conditions. Upon the onset of relax-
ation, the materials’ performance may be different from
INTRODUCTION what they were intended. When a building is subjected to
For the last two decades, carbon fiber-reinforced polymer a fire, reinforced concrete beams are in a stress relaxation
(CFRP) sheets have been employed to strengthen existing state caused by the out-of-plane displacement of adjacent
structural members. Although susceptible to elevated columns in tandem with temperature-induced loading (that
temperatures, these polymer-based materials are usable is, mechanical and thermal loads are applied simultane-
for buildings that are potentially vulnerable to fire hazards ously). Thermomechanical relaxation is, therefore, of interest
because insulation layers can protect the strengthened from structural integrity perspectives, including the case of
members and retard heat propagation into the CFRP system CFRP-strengthened beams. Unlike conventional relaxation,
for a certain period of time (fire rating). Williams et al.1 thermomechanical relaxation is not necessarily engaged
conducted a fire test with reinforced concrete T-beams with stabilized stresses because the functionality of a relaxed
strengthened with CFRP sheets in a furnace. Insulation material in thermal distress, such as with fire, is terminated
materials (thickness = 25 to 38 mm [1 to 1.5 in.]) were within a relatively short period of time. As reviewed earlier,
sprayed to protect the beams prior to testing. The maximum the temperature-dependent behavior of CFRP-strength-
furnace temperature was approximately 900°C (1652°F) ened concrete members alongside CFRP-concrete interface
at 2 hours and the corresponding temperature of the has been studied; however, their thermomechanical relax-
CFRP-concrete interface inside the insulation was less than ation responses have not been reported. Consequently, it is
200°C (392°F). Kodur and Yu2 proposed a design approach
for CFRP-strengthened members subjected to a fire. A rela- ACI Structural Journal, V. 115, No. 1, January 2018.
MS No. S-2017-088, doi: 10.14359/51701091, was received March 11, 2017, and
tionship between the members’ capacity and fire exposure reviewed under Institute publication policies. Copyright © 2018, American Concrete
Institute. All rights reserved, including the making of copies unless permission is
time was elaborated. Recommendations included that the obtained from the copyright proprietors. Pertinent discussion including author’s
approach used in current practice is conservative and equiv- closure, if any, will be published ten months from this journal’s date if the discussion
is received within four months of the paper’s print publication.

ACI Structural Journal/January 2018 259


unknown how CFRP systems perform when a strengthened
member is in thermomechanical relaxation and how the
degraded systems alter the efficacy of structural strength-
ening. This paper experimentally investigates the mecha-
nism of thermomechanical relaxation for CFRP-strength-
ened beams and analytically characterizes their behavior in
various conditions.

RESEARCH SIGNIFICANCE
Stress relaxation is considered important for structural
members consisting of multiple built-up components such
as externally bonded CFRP sheets that require appropriate
stress transfer to increase the load-carrying capacity of a
strengthened beam. The performance of a viscoelastic poly-
meric adhesive linking the concrete substrate and CFRP is
susceptibly influenced by thermal distress, and synergetic
degradation is expected when mechanical load is sustained.
This crucial load-carrying mechanism for CFRP-strength-
ened beams has been undervalued and is not considered
in design because no guidelines have been available. The
present research aims to understand the time- and tempera-
ture-dependent relaxation of reinforced concrete beams
strengthened with CFRP sheets, which results in analytical
expressions that can be used in practice.

EXPERIMENTAL INVESTIGATION
A test program is conducted to examine the thermome-
chanical relaxation and residual behavior of CFRP-strength- Fig. 1—Beam details: (a) dimension; and (b) cross section.
ened beams in flexure. Summarized in the following sections (Note: Unit in mm; 1 mm = 0.0394 in.)
are the materials, specimen preparation, loading schemes, 95% at 20°C [68°F]), the beams were taken out for drying
and instrumentation applied. and surface preparation using a wire brush to improve the
bond between the CFRP and substrate. The blended epoxy
Materials was applied along the tensile soffit of the beam, followed
Concrete was mixed for a compressive strength of fc′ = by the impregnation of single-layer dry carbon fiber
20 MPa (2900 psi), close to the average 28-day cylinder fabric (100 mm [4 in.] wide by 900 mm [35 in.] long). In
strength of 19.4 MPa (2813 psi). As primary reinforcement, compliance with the manufacturer’s recommendation, the
Grade 60 No. 3 steel bars were employed (yield strength σy CFRP-strengthening system was cured at room temperature
= 414 MPa [60 ksi] and cross-sectional area As = 71 mm2 for at least 7 days before testing. To preclude the premature
[0.11 in.2]). Shear stirrups were A36 plain steel rods (6 mm end-peeling failure of the longitudinal sheet, CFRP U-wraps
[0.24 in.] in diameter with a yield strength of 250 MPa (200 mm [8 in.] by 400 mm [16 in.]) were additionally
[36 ksi]). The unidirectional CFRP sheets used require bonded (Fig. 1).
a wet-lay-up process (that is, dry fibers are impregnated
in a resin matrix), and have a nominal tensile strength of Test procedure
σfu = 3800 MPa (551 ksi) and a modulus of Ef = 227 GPa The strengthened beams were simply supported with
(32,900 ksi) based on an equivalent thickness of tf = a span length of 1000 mm (3.3 ft). To generate thermal
0.165 mm (0.0065 in.). The resin and hardener components loading, a silicon rubber heat pad (150 mm [6 in.] long
of a two-part structural epoxy were blended at a mass ratio of by 100 mm [4 in.] wide) was attached to the tensile soffit
3:1 until a homogeneous mixture was attained. The specified of the beam at midspan (Fig. 1) by tightening steel wires,
yield strength and modulus of the epoxy are σey = 54 MPa which is the most vulnerable location in forming a plastic
(7830 psi) and Ee = 3 GPa (440 ksi), respectively, including hinge when mechanically loaded. The pad is composed of
a glass transition temperature of Tg = 71°C (160°F) with a fiberglass reinforcement and perfluoroalkoxy lead wires for
thermal conductivity of ke = 0.21W/(m·K) (1.45 Btu·in./ heat generation. Five steady-state temperature categories
(h·ft2·°F)). were tested from 75 to 150°C (167 to 302°F) an interval
of 25°C (45°F). Because the thermal behavior of the adhe-
Beam details sive at temperatures below its glass transition temperature
Singly reinforced concrete beams were cast with dimen- (Tg = 71°C [160°F]) is invariant, temperatures lower than
sions of 100 mm (4 in.) wide by 165 mm deep (6.5 in.) by 75°C (167°F) were not taken into consideration except for a
1200 mm (48 in.) long, as illustrated in Fig. 1. After 28 days control temperature of 25°C (77°F). To identify the flexural
of curing in an environmental chamber (relative humidity = capacity of the strengthened section (Puc) under three-point

260 ACI Structural Journal/January 2018


Fig. 2—Test setup: (a) instrumentation; and (b) thermocouple reading. (Note: °F = °C(9/5) + 32.)

Table 1—Test specimens relaxation periods, all beams were loaded to failure at a rate
of 2 mm/min (0.08 in./min) while maintaining the sustained
Load, kN (kip)
Thermomechanical heat for residual strength investigations.
Beam ID relaxation time, min. Yielding Ultimate
B25-0 0 42.7 (9.6) 52.4 (11.8) Instrumentation
B75-30 30 38.9 (8.7) 50.5 (11.3) To measure the applied load and displacement of the
respective beams, a load cell and a linear potentiometer were
B100-30 30 38.9 (8.7) 51.8 (11.6)
placed at midspan, as shown in Fig. 2(a). Displacement-type
B125-30 30 37.8 (8.4) 49.7 (11.1) strain transducers (PI gauges) were also mounted onto one
B150-30 30 37.3 (8.3) 47.6 (10.7) side of each beam to monitor tension and compression
B150-45 45 36.6 (8.2) 47.5 (10.6) strains. The progression of heat toward the beam concrete
was quantified by an infrared camera (Fig. 2(a)). The camera,
B150-60 60 35.8 (8.0) 47.3 (10.6)
built upon multi-spectral dynamic imaging (MSX) technol-
B150-75 75 35.2 (7.9) 43.9 (9.8) ogies, has a 19,200 pixel resolution with a thermal sensi-
tivity of less than 0.06°C (0.11°F) and a frame rate of 9 Hz.
bending (Fig. 2(a)), one beam was monotonically loaded Thermocouple wires (accuracy = ±2.2°C [±3.9°F]) were
to failure at 25°C (77°F). Other beams were preheated at positioned in between the CFRP and the heat pad to measure
midspan (Fig. 2(b)) to achieve the predefined temperatures, the applied temperatures (Fig. 2(b)). The thermomechanical
and were subjected to a load of 60%Puc for a default period response of the CFRP-concrete interface was studied by six
of 30 minutes during which the stroke of the universal strain gauges bonded outside the heating region (Fig. 1).
testing machine was fixed for a relaxation test. To examine
the effects of thermomechanical relaxation time, additional TEST RESULTS
loading periods up to 75 minutes were adopted to the beams The results of a two-phase experimental investigation
exposed to 150°C (302°F). These test schemes were summa- (thermomechanical relaxation and residual behavior) are
rized in Table 1 with the beams’ identification. For example, discussed with a focus on heat propagation, interfacial char-
B125-30 indicates a beam loaded at a temperature of 125°C acteristics, flexural capacity, and failure modes.
(257°F) for 30 minutes. The preceding thermomechanical
relaxation periods spanned over 40% of a 3-hour fire rating, Thermomechanical loading
which are sufficient from practice standpoints because Figure 3(a) exhibits the time-history plot of the thermo-
the restrained boundary condition of a flexural member is mechanical and residual tests for the beams subjected to
typically sustained for approximately 30 to 60 minutes in 150°C (302°F). The load was precipitously increased up to
a building under the occurrence of a fire.6,7 It should be approximately 30 kN (6.7 kip), at which the machine stroke
noted that the temperatures delineated herein are intended to was held for examining the beams’ relaxation behavior. The
represent those inside insulation layers enclosing structural applied loads gradually decayed until a residual capacity
beams when a fire takes place (that is, the relevant tempera- test was conducted. The thermal loading associated with
ture is the one at the interface level, instead of temperatures this mechanical testing scheme was reasonably maintained,
outside the structural member). Because the objective of as shown in the thermocouple measurement (Fig. 2(b)).
the present experimental program was to evaluate the influ- During the relaxation period, the mechanical loading rate
ence of thermomechanical relaxation on the behavior of did not markedly vary (Fig. 3(b)), including an alteration
CFRP-strengthened beams, rather than to study an increase range of less than 0.2 kN/s (0.05 kip/s). This indicates that
in load-carrying capacity after strengthening, unstrength- the sustained load relaxed in a consistent manner, which
ened beams were not tested. Upon completion of the planned confirms the adequacy of the present test setup.

ACI Structural Journal/January 2018 261


Fig. 3—Load-time history of beams: (a) response at 150°C (302°F); and (b) thermomechanical relaxation loading at 150°C
(302°F) for 45 minutes. (Note: 1 kN = 0.225 kip.)

Fig. 4—Heat propagation based on infrared thermal imaging: (a) Location 1; (b) Location 2; and (c) Location 3. (Note: °F
= °C(9/5) + 32.)
Heat progression near the heat pad. For instance, reductions of 9.7%, 23.4%,
The interface-level heat that propagated up the concrete is 40.3%, and 32.9% were recorded at the location of Gauge 4
provided in Fig. 4, where the measured temperatures (contours during the 30-minute relaxation period for the beams
in the infrared images) were compared under thermal loading exposed to 75, 100, 125, and 150°C (167, 212, 257, and
for 30 minutes. At Location 1 (Fig. 4(a)), the heat progres- 302°F), respectively. Furthermore, heat progression along
sion rate of the beam subjected to 150°C (302°F) was more the horizontal bond line was responsible for weakening the
rapid than that of the other beams. The progression rates of CFRP-concrete interface, which will be elaborated in the
the beams at Location 2 were virtually identical up to approx- subsequent section.
imately 20 minutes, regardless of interfacial temperature, after
which bifurcations were noticed. This observation implies Residual behavior
that a so-called latent heat effect exists in the concrete above The full loading history of the thermomechanically loaded
the CFRP bond-line. In other words, the applied heat energy beams at midspan is provided in Fig. 6, including residual
at the CFRP-concrete interface level does not immediately load-displacement responses (the B25-0 beam’s behavior was
change the temperature of nearby concrete until the energy similar to others; hence, it was not plotted for clarity). The
is sufficiently accumulated, which is a function of distance U-wrap anchors prevented end peeling of the longitudinal
from the interface. The far-field temperature response at Loca- CFRP, which led to the ductile behavior of the strengthened
tion 3 was marginal; nevertheless, the interfacial heat kept beams. It should be noted that the strains provided by the PI
conducting with time, as shown in Fig. 4(c). gauges were not useable on most occasions because the adhe-
sive employed for bonding the gauges to the concrete surface
Interfacial response during relaxation was affected by the heat propagation. For the cases subjected
Figure 5 shows the time-dependent strain of the CFRP to the relaxation period of 30 minutes (Fig. 6(a)), there was no
sheets. When the relaxation loading began, CFRP strains marked difference in flexural response up to approximately
close to the heat pad were generally higher than the strains 30 kN (6.7 kip). Thereafter, some divergence was noticed
of other locations. This is attributable to a curvature gradient owing to the adjustment of displacement along with relaxation
along the loading span (that is, the curvature at midspan of testing (further discussions will follow). Despite the likeness
the beam developed more rapidly than the curvature in the of load-displacement curves, the beams’ yield and ultimate
shear spans). With an increase in relaxation time, the strains loads were influenced by elevated temperatures, as enumer-
consistently dwindled. The degree of strain drop appeared ated in Table 1. With an increase in temperature from 75 to
to be related to elevated temperatures, particularly the ones 150°C (167 to 302°F), the yield and ultimate loads decreased

262 ACI Structural Journal/January 2018


Fig. 5—CFRP strain during thermomechanical relaxation: (a) B75-30; (b) B100-30; (c) B125-30; and (d) 150-30. (Note: °F
= °C(9/5) + 32.)

Fig. 6—Load-displacement at midspan: (a) elevated temperatures; and (b) variable thermomechanical relaxation periods.
(Note: 1 kN = 0.225 kip; 1 mm = 0.0394 in.)
by 4.1% and 5.7%, respectively. These were attributable to the relaxation time at 150°C (302°F), the decrease ratio did not
deteriorated CFRP-concrete interface, which lowered the effi- change considerably from 30 to 60 minutes, whereas a rapid
cacy of strengthening (that is to say, the extent of composite reduction in the ultimate load was observed during the expo-
action between the CFRP and concrete beam was dimin- sure period of 75 minutes. It is thus stated that thermomechan-
ished). Figure 6(b) exhibits the flexural behavior of the beams ical relaxation is crucial for CFRP-strengthened beams, and
exposed to 150°C (302°F) for variable relaxation periods. should be taken into consideration in design, wherever neces-
As before, the load-displacement responses were generally sary, with an emphasis on the degradation of the CFRP-con-
similar; however, their load-carrying capacities were a func- crete interface that controls the beams’ performance.
tion of thermal exposure time: decreases of 5.6% and 7.8% Figure 8(a) demonstrates CFRP-strain profiles for the
of the yield and ultimate loads of the beams were recorded, B25-0 beam at an interval of 25% of the ultimate load (Pu).
respectively, when the thermomechanical relaxation period Following the increasing flexural load, the profiles gradu-
changed from 30 to 75 minutes (Table 1). A summary of the ally developed. At failure of the beam by concrete crushing
temperature effects on altering the beams’ residual capacity is (100%Pu), erratic strain gradients were observed, which
shown in Fig. 7. Except for the minor discrepancy in B75-30 denote that the longitudinal CFRP partially debonded. The
and B100-30 that failed at 50.5 kN (11.3 kip) and 51.8 kN U-wrap anchors preserved the interfacial integrity; conse-
(11.6 kip), respectively, the decrease ratio of the ultimate load quently, the strains at Gauges 1 and 6 in the vicinity of the
almost linearly varied with elevated temperatures. In terms of anchors (G1 and G6, respectively) were lower than others.

ACI Structural Journal/January 2018 263


The propensity for strain development, however, changed Possible reasons for that are 1) the empirical ACI equation
when the beams were exposed to thermomechanical loading, was calibrated without due consideration about the use of
as shown in Fig. 8(b). The B150-75 beam’s strain profiles U-wrap anchors; and 2) the present CFRP properties were
were similar to those of B25-0 until a load level of 50%Pu outside the calibration range. In either case, an improvement
was accomplished (the flexural load promptly increased in in the ACI expression appears to be necessary. The CFRP-
the nascent loading stage [Fig. 3(a)]; therefore, the effect strain variation caused by the thermomechanical relaxation
of elevated temperatures was not yet apparent). Passing is given in Fig. 8(c) and (d). The B75-30 beam revealed a
through the relaxation period, the strain profile became strain retraction from a load of 33.3 kN (7.5 kip) down to
asymmetric (75%Pu) and, eventually, the beam failed with a 26.8 kN (6.0 kip), accompanying a 24.7% strain drop in
rapid increase in strain on the G1 gauge side (100%Pu). This strain Gauge 6 (G6 in Fig. 8(c)). The retraction of the CFRP
fact signifies that the CFRP-concrete interface weakened strains was pronounced with the increased temperature and
during thermomechanical relaxation, and was disintegrated relaxation time, as shown in Fig. 8(d), which impacted the
by the residual loading with interfacial slippage between the integrity of the CFRP-concrete interface and, thereby, the
CFRP and concrete. It is worthwhile to note that debonding effectiveness of CFRP-strengthening.
strains recorded in these beams were lower than the effec-
tive strain determined by ACI 440.2R-088: εfd = 0.0095. Failure mode
The representative failure modes of the test beams are
provided in Fig. 9. The interfacial integrity of B25-0 loaded
without thermomechanical relaxation was preserved until the
concrete crushed (Fig. 9(a)). In conjunction with the insignif-
icant development of diagonal tension cracks, intermediate
crack-induced CFRP-debonding (known as IC-debonding)
was not observed along the loading span. For the beam
subjected to thermomechanical loading (Fig. 9(b)), the
CFRP-concrete interface was deteriorated during the relax-
ation period; subsequently, minor cracks formed upon the
residual loading. The cracks initiated at the junction of the
longitudinal sheet and the U-wrap anchor were connected to
the primary diagonal tension crack, which progressed toward
the loading point where concrete crushing eventually took
place. The tendency of the interface deterioration was obvi-
Fig. 7—Variation of residual capacity. (Note: °F = °C(9/5)
ously augmented with the increased temperature and relaxation
+ 32.)

Fig. 8—Load-CFRP strain up to failure: (a) profile of B25-0; (b) profile of B150-75; (c) development of B75-30; and (d) relax-
ation effect. (Note: 1 kN = 0.225 kip; 1 mm = 0.0394 in.)

264 ACI Structural Journal/January 2018


Fig. 9—Failure mode: (a) B25-0; (b) B100-30; and (c) B150-75.
period, as shown in Fig. 9(c). In accordance with these failure Taking an exponential form for the creep-strain rate of
modes, it is confirmed that the thermal loading progressed Eq. (3) (dεfc(t,T)/dt = σf(t,T)m) along with the CFRP’s consti-
along the interface (although no physical CFRP-debonding tutive relationship [εfe(t,T) = σf(t,T)/ Ef(T)] yields
was associated during the relaxation period) and degraded the
interfacial characteristics that brought about strain localization d σ f (t , T )
adjacent to the U-wrap anchors. The presence of the anchors = − dt (4)
σ f (t , T ) m
precluded end-peeling of the CFRP; accordingly, the load-car-
rying capacity of the beams was maintained with a maximum where m is a relaxation constant. Integrating Eq. (4)
variation margin of 7.8%, in spite of the detrimental thermo-
(1− m)
mechanical effects, as discussed earlier. σ f (t , T )
= −t + C (5)
1− m
ANALYTICAL MODELING
An analytical model is developed based on conventional where C is the integration constant to be solved by the CFRP
relaxation theory9 to predict the thermomechanical relaxation stress immediately before thermomechanical relaxation
of CFRP-strengthened beams. Upon calibrating the model loading (σf(0,T) at t = 0 minutes). Rearranging Eq. (5) gives
using test data, parametric investigations are conducted and
design recommendations are proposed. σf(t,T) = [(m – 1)t + σf(0,T)(1 –m)]1/(1 –m) (6)

Development
The time-dependent strain of the CFRP at temperature T Calibration and assessment
[εf(t,T)] is expressed as To calibrate the relaxation constant m, a transformed
section of the strengthened beam at midspan may be used
εf(t,T) = εfe(t,T) + εfc(t,T) (1)
n f (T ) M (t , T ) y f (T )
σ f (t , T ) = (7)
where εfe(t,T) and εfc(t,T) are, respectively, the elastic and I t (T )
creep components of the CFRP strain (hereafter, t and T indi-
cate time- and temperature-dependent properties, respec- where nf(T) is the CFRP modular ratio for the transformed
tively). According to previous research,10 the tempera- section at temperature T; M(t,T) is the time-dependent
ture-dependent modular ratio [nf(T)] between typical moment applied; yf(T) is the distance from the neutral axis
CFRP and concrete (Ef(T) and Ec(T), respectively) within a to the CFRP level; and It(T) is the moment of inertia of the
temperature range from 25 to 200°C (77 to 392°F) may be transformed section. It is assumed that, when determining
obtained by (T in Celsius) the moment of inertia It(T), the modulus of steel reinforce-
ment is virtually temperature-independent below 150°C
E f (T ) E fo exp( −0.0009T ) (302°F).12 Substituting Eq. (6) into Eq. (7) with the boundary
n f (T ) = = (2)
Ec (T ) Eco ( −0.001282T + 1.025641) condition employed in the test program (Fig. 1)

where Efo and Eco are the initial CFRP modulus and concrete P(t,T) = kΔ for 0 ≤ t ≤ tR (8a)
modulus (Eco = 57,000√fc′ in psi [4730√fc′in MPa11]), respec-
tively. In a thermomechanical relaxation condition of the 4σ f (t , T ) I t (t , T )
P (t , T ) = for tR < t ≤ tl (8b)
CFRP-strengthened beam (that is, deformation of the CFRP n f (t , T ) y f (t , T ) L
remains unchanged), the strain rate becomes zero
where P(t,T) is the load associated with thermomechanical
d ε f e (t , T ) d ε f c (t , T ) relaxation at time t in min. (Fig. 3(a)); k is the stiffness of
0= + (3) the strengthened beam; Δ is the displacement of the beam
dt dt

ACI Structural Journal/January 2018 265


before the relaxation begins at time tR; and tl is the time when nificant discrepancy between the tested and predicted load
relaxation ends. The experimental load variations of the variations (for instance, a margin of 6% at 30 minutes) was
beams subjected to a thermomechanical relaxation period due to the use of the regression-based constant (Eq. (9)). All
of 30 minutes (Fig. 3(a)) were substituted into Eq. (6) and tested and predicted thermomechanical relaxation responses
(8b) in tandem with the temperature-dependent variables. during the 30-minute period (tR < t ≤ tl) at an interval of 1
The unknown temperature-dependent constant m was then second were compared in Fig. 11(b), including an average
determined by a generic algorithm. This computational absolute difference of 4.6%. Overall, the model prediction
technique is useful to solve a complex problem based on a was in agreement with the experimentally-attained relax-
mutation-and-selection approach with randomly generated ation data, and can be useable for parametric investigations
data samples. Further details on the solution algorithm and detailed in the ensuing section.
its theoretical background are available elsewhere.13Figure
10 summarizes the average values of the constant from the Parametric study
initiation of the relaxation (tR) to its end (tl) at the respective During a thermomechanical relaxation period (tR < t ≤ tl),
temperatures. As the temperature rose, the m values gradually the amount of relaxation in CFRP is determined by
increased, which resulted in the following best-fit equation (T
in Celsius) (σ f (0, T ) − σ f (t , T ))
λR = (10)
σ fu
m = 0.1406ln(T) – 0.4695 (9)
where λR is the thermomechanical relaxation factor. Figure 12
With the calibrated constant, Eq. (8a) and (8b) were exhibits the variation of the relaxation factor with several param-
appraised with the experimental data, as shown in Fig. 11. eters that typically influence the behavior of CFRP-strengthened
It should be noted that the load at the initiation of relaxation beams. For comparison, unless otherwise stated, the present test
(Eq. (8a)) was taken from the test data because the beams’ configuration depicted in Fig. 1 alongside the aforementioned
response before and after the relaxation (initial loading and material properties was the default within a temperature range
residual loading, respectively) was outside the interest of the from 75 to 150°C (167 to 302°F) for a 30-minute relaxation
present study. Figure 11(a) assesses the predicted thermome- period. As illustrated in Fig. 12(a), elevated temperatures were
chanical relaxation behavior of the beam at a temperature of crucial in augmenting the degree of thermomechanical relax-
75°C (167°F) against its experimental counterpart. The insig- ation. The relaxation factor associated with 150°C (302°F) was
174% higher than the factor with 75°C (167°F) at 30 minutes
(Fig. 12(a), inset). This is explained by the increased mobility
of polymeric chains in the epoxy resin with temperature.14
Because the relaxation was most noticeable at 150°C (302°F),
the predicted results of the parametric study at this tempera-
ture are only discussed hereafter. The effects of CFRP modulus
are provided in Fig. 12(b). The higher the CFRP modulus, the
more the relaxation amount was resulted. The reason is that
the CFRP stress σf(0,T) of the transformed section increased
with the increased CFRP modulus (Eq. (7)). It is important
to note that this response trend in the strengthened beam does
not necessarily mean CFRP itself is susceptible to relaxation.
The thermomechanical relaxation of the strengthened beam
was marginally influenced by concrete strength (Fig. 12(c)),
whose contribution to the transformed properties of the beam
Fig. 10—Determination of thermomechanical relaxation was negligible. When a CFRP stress level at time tR increased,
constant. (Note: °F = °C(9/5) + 32.)

Fig. 11—Assessment of predictive model: (a) comparison at 75°C (167°F); and (b) comparison at all temperatures during
thermomechanical relaxation periods. (Note: °F = °C(9/5) + 32; 1 kN = 0.225 kip.)

266 ACI Structural Journal/January 2018


Fig. 12—Effects of parameters on thermomechanical relaxation: (a) temperature; (b) CFRP modulus; (c) concrete strength;
and (d) initial stress level (fraction of CFRP strength). (Note: °F = °C(9/5) + 32; 1 kN = 0.225 kip; 1 MPa = 145 psi.)

Table 2—Regression constants


Thermomechanical relaxation time
Constant 10 min 20 min 30 min Application range
c1 0.0007 0.0015 0.0023
c2 –0.0001 –0.0005 –0.0010 75°C ≤ T ≤ 150°C
100 GPa ≤ Ef ≤ 300 GPa
c3 0.0004 0.0007 0.0010
20 MPa ≤ fc′ ≤ 40 MPa
c4 0.0060 0.012 0.0170 0.1ffu ≤ σf(0,T) ≤ 0.5ffu
c5 –0.1500 –0.2900 –0.4320

Notes: °F = °C(9/5) + 32; 1 MPa = 145 psi.

the strengthened beam experienced more relaxation, as shown


in Fig. 12(d). At a stress level of 0.5σfu, the relaxation factor
increased up to 4.6%, which is approximately 50% greater than
the factor at a stress level of 0.1σfu. This fact signifies that the
magnitude of mechanical load sustained prior to relaxation is Fig. 13—Assessment of regression equation.
influential on the relaxation of CFRP-strengthened beams. SUMMARY AND CONCLUSIONS
The aforementioned parametric study results were inte- This paper has discussed the thermomechanical relaxation
grated by the multiple regression method so that a unified and residual behavior of reinforced concrete beams strength-
expression was acquired at a specific relaxation period (for ened with CFRP sheets. Test parameters encompassed
convenience, the following units are used: Ef in GPa, fc′ in temperatures and relaxation periods. An infrared imaging
MPa, σf(0,T) in MPa, and T in Celsius) technique was employed to examine the heat conduction
of the strengthened beams across the CFRP-concrete inter-
λR(%) = c1Ef + c2fc′ + c3σf(0,T) + c4T + c5 (11) face. The development of CFRP strains in conjunction
with various failure modes of the beams was emphasized.
where ci is a regression constant, available in Table 2. Figure 13 Analytical modeling based on the temperature-dependent
evaluates the accuracy of the regression equation using the test characteristics of the CFRP system was conducted to predict
data, both of which agreed with an average absolute difference the degree of thermomechanical relaxation, including a
of 2.1%. calibrated expression for relaxation constant, and was used
for parametric investigations. A thermomechanical relax-
ation factor was proposed based on the multiple regression

ACI Structural Journal/January 2018 267


method, which can assist practitioners involved in CFRP- ACKNOWLEDGMENTS
based rehabilitation for building structures. The writers gratefully acknowledge financial support from Saudi Arabian
Cultural Mission and the U.S. Department of Transportation through the
• During thermomechanical relaxation, a latent heat effect Mountain-Plains Consortium. Proprietary information such as manufac-
was noticed above the CFRP-concrete interface, which turers and product names was not shown to avoid commercialism. Tech-
retarded temperature progression into the concrete nical contents presented in this paper are based on the opinion of the writers
and do not necessarily represent that of others.
beam. Thermal conduction along the CFRP weakened
the strength of the interface.
REFERENCES
• Thermomechanical relaxation caused partial CFRP- 1. Williams, B.; Kodur, V.; Green, M. F.; and Bisby, L., “Fire Endurance
debonding when mechanically loaded in residual testing of Fiber-Reinforced Polymer Strengthened Concrete T-Beams,” ACI Struc-
and the measured debonding strains were lower than the tural Journal, V. 105, No. 1, Jan.-Feb. 2008, pp. 60-67.
2. Kodur, V. K. R., and Yu, B., “Rational Approach for Evaluating Fire
effective strain calculated by ACI 440.2R-08. The extent Resistance of FRP-Strengthened Concrete Beams,” Journal of Compos-
of CFRP-strain retraction representing the integrity of the ites for Construction, ASCE, V. 20, No. 6, 2016, 04016041. doi: 10.1061/
CFRP-concrete interface (composite action) was affected (ASCE)CC.1943-5614.0000697
3. Hu, W., Polymer Physics: A Molecular Approach, Springer, Berlin,
by elevated temperatures and relaxation time. Germany, 2013.
• The residual capacity of the strengthened beams was 4. Gamage, J. C. P. H.; Al-Mahaidi, R.; and Wong, M. B., “Bond Charac-
reduced by thermomechanical relaxation, which also teristics of CFRP Plated Concrete Members under Elevated Temperatures,”
Composite Structures, V. 75, No. 1-4, 2006, pp. 199-205. doi: 10.1016/j.
influenced the load-displacement behavior and interfa- compstruct.2006.04.068
cial cracking patterns of the beams. The trend of capacity 5. Leone, M.; Matthys, S.; and Aiello, M. A., “Effect of Elevated Service
reduction was proportional to the degree of temperatures. Temperature on Bond between FRP EBR Systems and Concrete,” Compos-
ites. Part B, Engineering, V. 40, No. 1, 2009, pp. 85-93. doi: 10.1016/j.
• Unlike concrete strength, the elastic modulus of CFRP compositesb.2008.06.004
and the magnitude of mechanical load augmented the 6. Tan, K.-H.; Toh, W.-S.; Huang, Z.-F.; and Phng, G. H., “Structural
thermomechanical relaxation of the strengthened beams. Responses of Restrained Steel Columns at Elevated Temperatures,” Engi-
neering Structures, V. 29, No. 8, 2007, pp. 1641-1652. doi: 10.1016/j.
The proposed relaxation factor is recommended to engstruct.2006.12.005
address the potential degradation of CFRP systems when 7. Sadaoui, A., and Khennane, A., “Effect of Transient Creep on the
subjected to a thermomechanical loading condition. Behaviour of Reinforced Concrete Columns in Fire,” Engineering Structures,
V. 31, No. 9, 2009, pp. 2203-2208. doi: 10.1016/j.engstruct.2009.04.005
8. ACI Committee 440, “Guide for the Design and Construction of Exter-
AUTHOR BIOS nally Bonded FRP Systems for Strengthening Concrete Structures (ACI
Yail J. Kim, FACI, is a Professor in the Department of Civil Engineering 440.2R-08),” American Concrete Institute, Farmington Hills, MI, 2008, 76 pp.
at the University of Colorado Denver, Denver, CO. He is the Chair of ACI 9. Kraus, H., Creep Analysis, John Wiley & Sons, New York, 1980.
Committee 345, Concrete Bridge Construction, Maintenance, and Repair, 10. Kim, Y. J.; Hmidan, A.; and Choi, K.-K., “Residual Behavior of
and ACI Subcommittee 440-I, FRP-Prestressed Concrete. He is a member of Shear-Repaired Concrete Beams Using CFRP Sheets Subjected to Elevated
ACI Committees 342, Evaluation of Concrete Bridges and Bridge Elements, High Temperatures,” Journal of Composites for Construction, ASCE, V. 16,
and 440, Fiber-Reinforced Polymer Reinforcement; and a member of Joint No. 3, 2012, pp. 253-264. doi: 10.1061/(ASCE)CC.1943-5614.0000244
ACI-ASCE Committee 343, Concrete Bridge Design. His research inter- 11. ACI Committee 318, “Building Code Requirements for Structural
ests include advanced composite materials for structural rehabilitation, Concrete (ACI 318-14) and Commentary,” American Concrete Institute,
complex systems, uncertainty quantification, and science-based structural Farmington Hills, MI, 2014, 520 pp.
engineering including statistical, interfacial, and quantum physics. 12. ACI Committee 216, “Guide for Determining the Fire Endurance
of Concrete Elements (ACI 216R-89) (Reapproved 2001),” American
Abdulaziz Alqurashi was a Graduate Student in the Department of Civil Concrete Institute, Farmington Hills, MI, 1989, 48 pp.
Engineering at the University of Colorado Denver, where he received his 13. Mitchell, M., An Introduction to Genetic Algorithms, MIT Press,
MS in civil engineering in 2015. His research interests include the thermal Cambridge, MA. 1998.
behavior of structural concrete with advanced materials. 14. Casalini, R.; Fioretto, D.; Livi, A.; Lucchesi, M.; and Rolla, P. A.,
“Influence of the Glass Transition on the Secondary Relaxation of an Epoxy
Resin,” Physical Review B: Condensed Matter and Materials Physics,
V. 56, No. 6, 1997, pp. 3016-3021. doi: 10.1103/PhysRevB.56.3016

268 ACI Structural Journal/January 2018


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 115-S22

Practicability of Large-Scale Reinforced Concrete Beams


Using Grade 80 Stirrups
by Jung-Yoon Lee, Jae-Hoon Lee, Do Hyung Lee, Seong-Jun Hong, and Ho-Young Kim

ACI 318-14 specifies that the yield strength of shear reinforcement easier concrete placement and, thus, an improved concrete
is limited to 420 MPa (60 ksi) for shear-strength calculation of rein- quality.
forced concrete beams to control failure mode and diagonal crack Application of the high-strength steel in reinforced concrete
width for such members. However, in the case where high-strength structures requires various investigations regarding material
shear reinforcement is employed in such members, the amount of the
characteristics, member level performance (that is, flexure,
shear reinforcement can be reduced and wider spacing is feasible.
shear, bond, anchorage, axial compression, and seismic
This may lead to an easy placement of concrete and, thus, improved
durability. Thus, the use of high-strength shear reinforcement can performance). In particular, because primary national codes
be particularly effective for structures having large-scale members for reinforced concrete structures (ACI 318-14,1 EU-04,2 and
such as skyscrapers, long-span bridges, and nuclear power plants. CSA-043) are mostly based on empirical design formulas,
To investigate the practicability of the high-strength shear rein- experimental evidence is required for the performance evalu-
forcement, shear tests are carried out for large-scale reinforced ation of such structures employing high-strength steel.
concrete beams using Grade 80 stirrups. For this purpose, a total For these reasons, much research has been dedicated to the
of 16 reinforced concrete beams are tested and categorized into development of high-strength steel over the last decade in
four groups according to the parameters: two levels of stirrup yield many countries. In the United States, a great deal of exper-
strength and four levels of cross section depth. Current test results imental research on shear, bond, and development length
reveal that all large-scale beams fail in shear tension, experiencing
has been conducted for the application of Grade 100 steel to
yielding of stirrups prior to concrete web crushing. Measured
reinforced concrete structures.4,5 In Japan, similar research
shear strength is greater than theoretical shear strength calculated
by ACI 318-14 in all members. However, the margin between the regarding the development of high-strength steel has been
two strengths is reduced as the cross section depth increases. The carried out, including an investigation of the characteristics
diagonal crack width is influenced by the stirrup yield strength and of axial compression, shear, and bond, as well as performing
spacing, and the cross section depth. The crack width increases as finite element analysis.6 In Taiwan, high-strength steel has
the cross section depth and stirrup spacing increase. been investigated through a national project: the Taiwan
New RC Project.7,8 In Korea, several research works have
Keywords: diagonal crack width; Grade 80 stirrups; large-scale reinforced been conducted on high-strength steel since 2006.9-12
concrete beams; shear strength; size effect.
In addition, works on the application of high-strength steel
to shear reinforcement have been conducted by Munikrishma
INTRODUCTION
et al.,13 Selim et al.,14 and Lee et al.10-12 These works revealed
Many studies on high-strength and high-performance
that, even for members in which Grade 80 high-strength stir-
concrete have been extensively carried out over the last
rups are employed, the stirrups reached yield strain before
30 years. Subsequently, high-strength concrete of over
maximum strength was attained, resulting in shear tension
80 MPa (11.6 ksi) has been used in design practice. However,
failure. Also, the average diagonal crack width was less than
contrary to high-strength concrete, a few studies have been
0.41 mm (0.016 in.), which is the recommended crack width
conducted with respect to the employment of high-strength
according to ACI 224R-01.15 These results imply that the
steel in reinforced concrete structures. The use of high-
use of Grade 80 high-strength steel as stirrups is promising.
strength steel in reinforced concrete structures can have
However, in the above works, the members are somewhat
beneficial effects on cost and constructability. Specifically,
limited in cross section size because most of the reinforced
special structures such as nuclear power plants, skyscrapers,
concrete beams are not larger than 600 mm (23.6 in.) in
long-span bridges, and liquid nitrogen gas (LNG) storage
section depth. Consequently, the shear behavior on large-
tanks require very large amounts of steel. For example,
scale beams in which high-strength stirrups are employed
the total amount of steel for the construction of a nuclear
needs to be evaluated, considering that such members are, in
power plant containment vessel is approximately 70,000 to
general, used in high-rise buildings, nuclear structures, and
100,000 tons. In addition, large-diameter steel reinforce-
other special structures such as long-span bridges. Nonethe-
ments of over D35 (No. 11) are employed in reinforced
concrete skyscrapers over 150 m (492 ft) in height. The
use of high-strength steel in such structures could lead to a
ACI Structural Journal, V. 115, No. 1, January 2018.
reduction of the total steel amount and, hence, a reduction MS No. S-2017-089.R1, doi: 10.14359/51701092, was received March 26, 2017,
and reviewed under Institute publication policies. Copyright © 2018, American
in cost. Moreover, the use of high-strength steel may result Concrete Institute. All rights reserved, including the making of copies unless
in a wider arrangement of reinforcement spacing, leading to permission is obtained from the copyright proprietors. Pertinent discussion including
author’s closure, if any, will be published ten months from this journal’s date if the
discussion is received within four months of the paper’s print publication.

ACI Structural Journal/January 2018 269


Table 1—Details of specimens and material properties
Longitudinal tensile reinforcement Stirrup
Specimen fc′, MPa d, mm h, mm b, mm ρl, % fyl, MPa s, mm ρv, % fyt, MPa
BS1-1 — — —

Group BS1-2 130 0.219 459


43.6 501.5 600 500 1.61 695
BS1 BS1-3 180 0.158 674
BS1-4 130 0.219 674
BS2-1 — — —

Group BS2-2 40.0 130 0.219 459


751.5 850 500 1.61 695
BS2 BS2-3 180 0.158 674
BS2-4 46.2 130 0.219 674
BS3-1 — — —

Group BS3-2 130 0.219 459


48.5 1001.5 1100 500 1.61 695
BS3 BS3-3 180 0.158 674
BS3-4 130 0.219 674
BS4-1 — — —

Group BS4-2 130 0.219 459


54.8 1251.5 1350 500 1.61 695
BS4 BS4-3 180 0.158 674
BS4-4 130 0.219 674

Notes: 1 mm = 0.0394 in.; 1 MPa = 145 psi.

less, very few studies have been carried out for the applica- In the present study, experimental work was conducted on
tion of Grade 80 high-strength stirrups to those members. large-scale reinforced concrete beams with Grade 80 high-
Among such studies, works by Kani,16 Zararis and Papa- strength stirrups. The yielding of high-strength stirrups,
dakis,17 Bentz,18 and Bažant19 demonstrated that the shear diagonal crack width, and size effect are observed and, thus,
behavior of large-scale reinforced concrete beams was influ- the applicability of high-strength stirrups to large-scale rein-
enced by the size effect, and the size effect was closely asso- forced concrete beams is evaluated.
ciated with the effective depth of a section and shear span-
depth ratio. Qiang and Bažant20 investigated the size effect RESEARCH SIGNIFICANCE
of reinforced concrete beams with stirrups, and concluded Most studies on the applicability of high-strength shear
that, while stirrups were able to mitigate the size effect, reinforcement focus on structural members with relatively
they were not able to suppress it completely. Much research small section dimensions. However, high-strength shear
has been dedicated to diagonal crack width for reinforced reinforcement is generally employed in skyscrapers or large-
concrete beams in the last few decades. Subsequently, the scale structures rather than in mid- or low-rise structures.
influencing parameters and occurrence mechanism on the Therefore, investigation on the applicability of high-strength
crack have been investigated.21-25 Previous research has indi- shear reinforcement to large-scale members is important
cated that the spacing of the crack is reduced by an increase and necessary. When high-strength shear reinforcement
of shear reinforcement ratio, and the diagonal crack width is is employed in a large-scale member, the total amount of
increased by wider stirrup spacing. Lee et al.11,12 performed reinforcement can be reduced and, thus, a wider spacing
a test on the shear behavior of reinforced concrete beams between the reinforcements is possible. This can result in
with high-strength stirrups. They demonstrated that the an improved quality of concrete due to easy placement of
diagonal crack width is not proportional to the yield strength concrete. However, a detrimental effect may also occur,
of stirrups but is influenced by a number of stirrups (ρv fyt). whereby diagonal crack width can be increased in large-
From a test on large-scale beams, Shoya et al.26 showed that scale members with high-strength shear reinforcement, and
shear crack width was increased as the cross section of a the shear strength provided by the high-strength shear rein-
member increased. However, the aforementioned studies forcement can vary due to the size effect. To investigate this
are not sufficient enough to evaluate the size effect of rein- phenomenon, a total of 16 large-scale reinforced concrete
forced concrete beams. This is because most of the afore- beams employing Grade 80 high-strength stirrups are tested
mentioned studies are limited to either large-scale members in the present study. The test results are expected to provide
with normal-strength stirrups or relatively small beams with insight into the shear behavior of such members, as well
section depths of less than or equal to 600 mm (23.6 in.). as useful information on the feasibility of employing these
Therefore, research on the shear behavior of large-scale high-strength stirrups in such members.
beams of not less than 1000 mm (39.4 in.) in section depth
with Grade 80 high-strength stirrups is urgently needed.

270 ACI Structural Journal/January 2018


TEST SPECIMEN AND PROCEDURE depth ratio is designed to be 2.5 in all specimens, being shear
A total of 16 large-scale reinforced concrete beams were critical. Also, the tensile reinforcement ratio is designed to
constructed to evaluate the applicability of high-strength be 1.61%, which is identical in all specimens. The inten-
stirrups to the specimens. Details of the beam specimens and tion for the identical ratio is to ensure the same influence
material properties are summarized in Table 1. The primary of the tensile reinforcement on size effect. The ratio is set
parameter of the specimens is the cross section size. Four as relatively high to prevent flexure failure prior to shear
different section sizes are employed in this study: 500 x failure. Meanwhile, it is worth noting that longitudinal skin
600 mm (19.7 x 23.6 in.), 500 x 850 mm (19.7 x 33.5 in.), reinforcements are not arranged in the specimens of 500
500 x 1100 mm (19.7 x 43.3 in.), and 500 x 1350 mm (19.7 x 1100 mm (19.7 x 43.3 in.) and 500 x 1350 mm (19.7 x
x 53.1 in.), as shown in Fig. 1 and 2. For the stirrups, two 53.1 in.), although ACI 318-141 specifies that the longitu-
different types of steel, Grade 60 and Grade 80, are used in dinal skin reinforcement needs to be uniformly distributed
the fabrication of the specimens. Tensile tests on both types along both side faces of a beam in case where section depth
of steel indicate that the actual yield strengths of Grade 60 of the beam exceeds 914.4 mm (36 in.). This is to evaluate
and Grade 80 steel are 459 and 674 MPa (66.6 and 97.8 ksi), directly the influence of high-strength stirrups on crack
respectively, as shown in Table 1 and Fig. 3. The actual yield width for specimens with different section depths.
strength of Grade 80 steel used for longitudinal reinforce- Test specimens are categorized into four different groups
ment of No. 11 bar is 695 MPa (100.8 ksi). The shear span- (BS1, BS2, BS3, and BS4) depending on the cross section
sizes, and each group is composed of four specimens
in which the primary parameters are spacing and yield
strength of stirrup. The specimens numbered as 1- in each
group have no stirrups, while those numbered with 2- have
Grade 60 stirrups with 130 mm (5.12 in.) spacing, and those
numbered with 3- have Grade 80 stirrups with 180 mm
(7.09 in.) spacing. This arrangement leads to an identical
stirrup amount (ρv fyt) of 1.0 MPa (145 psi) in both speci-
mens. For the specimens numbered with 4-, Grade 80 stirrups
with 130 mm (5.12 in.) spacing are employed. The purpose
of these overall arrangements is to compare and evaluate
the effects of high strength stirrups on shear strength and
diagonal crack width in large-scale beams. Meanwhile, the
compressive strength of concrete is designed to be 45 MPa
(6635 psi), and the strength given in Table 1 represents the
actual strength at the date of the test.
All specimens are designed in accordance with the shear
equations given in ACI 318-141 (Eq. 22.5.1.1, 22.5.6.1(a),
and 22.5.10.5.3) and in accordance with provisions regarding
Fig. 1—Test setup and measurement.

Fig. 2—Overall dimensions of specimens and locations of linear variable differential transformers (LVDTs). (Note dimensions
in mm; 1 mm = 0.0394 in.)

ACI Structural Journal/January 2018 271


maximum and minimum stirrup ratio, and maximum stirrup in both shear-critical areas. For the diagonal crack width of
spacing in ACI 318-14.1 Particular emphasis is placed on the concrete struts in the shear critical areas, two schemes were
stirrup ratio, which is as low as possible to investigate the employed to measure the width precisely. Demac gauges and
size effect of large-scale beams. However, closely spaced crack magnifiers having a minimum scale division of 0.001
high-strength stirrups are arranged in both ends of the spec- and 0.01 mm (0.000039 and 0.00039 in.), respectively, were
imens to prevent support failure. In addition, to avoid local used in the test. All the tests were terminated when the load
concrete crushing or bearing failure at both loading points reached 90% of the maximum strength after the maximum
and supports, rectangular steel plates of 100 x 100 mm (3.94 strength was achieved.
x 3.94 in.) are placed at such a location.
Loading was applied using a universal testing machine STRUCTURAL BEHAVIOR
(UTM) of 5000 kN (1,124,045 lb) capacity with displace- Crack patterns
ment control, and the applied loading was recorded in a data Shear failures occurred in all specimens in which maximum
acquisition instrument (brand name; remove). Strain gauges strength was attained before the strain gauges attached to
were installed in both longitudinal tensile reinforcements the longitudinal tensile reinforcements reached yield strain.
and stirrups, as illustrated in Fig. 2 to investigate the yielding Representative specimens of the BS4 group are shown in
of both reinforcements, and, thus, the failure modes of the Fig. 4, where typical shear cracks are clearly shown. Flex-
specimens. The average deformation of the specimens was ural cracks were initially formed in the portion of pure flex-
estimated using 12 linear variable differential transformers ural moment between loading points. More flexural cracks
(LVDTs): three horizontal, two vertical, and one diagonal developed near loading points where the maximum flexural
moment and shear force occurred simultaneously. These flex-
ural cracks then changed to flexure-shear cracks as the load
was increased. When the load was near maximum strength,
shear cracks occurred in a web between the support and the
loading point. For the specimens without stirrups, diagonal
cracks having a large width occurred after the formation of
flexural cracks. The load was then slightly increased after
the formation of the diagonal cracks. However, the load
was then decreased immediately after additional diagonal
cracks developed. Many diagonal cracks in specimens with
stirrups were relatively greater than those in specimens
without stirrups. Diagonal crack width in specimens with
stirrups increased up to 2 mm (0.079 in.) when the maximum
strength was almost reached. The estimated diagonal crack
Fig. 3—Tensile stress-tensile strain relationship of steel
bars. (Note: 1 MPa = 145 psi.)

Fig. 4—Representative crack patterns after test for specimens of Group BS4.

272 ACI Structural Journal/January 2018


Fig. 5—Shear force-deflection relationship of each Group. (Note: Rectangular section dimensions are in mm; 1 mm = 0.039 in.;
1 kN = 224.8 lb.)
Table 2—Shear strength of specimens
width was in general proportional to the cross section size.
Vn-test, Vn-ACI, Vn-EC2,
Nonetheless, many cracks were almost constant. Specimen kN kN kN V /V V /V n-test n-ACI n-test n-EC2

BS1-1 524.9 293.1 303.8 1.79 1.73


Shear force-deflection relationship
The shear force-deflection relationship for all speci- Group BS1-2 833.5 544.6 567.1 1.53 1.47
mens of the four groups is shown in Fig. 5. In general, the BS1 BS1-3 727.7 559.8 600.8 1.30 1.21
maximum shear strength is proportional to the cross-section
BS1-4 807.1 662.3 832.8 1.22 0.97
depth. The shear strength developed in specimens without
stirrups started to decrease immediately after two to three BS2-1 424.9 422.8 411.0 1.01 1.03
diagonal cracks had formed. However, the deflection in Group BS2-2 1239.1 799.6 849.8 1.55 1.46
such specimens was not significant. Similar behavior was BS2 BS2-3 915.5 822.4 900.3 1.11 1.02
observed in specimens with stirrups where the shear strength
BS2-4 1130.4 1004.5 1247.9 1.13 0.91
was reduced abruptly once the maximum shear strength was
attained. Figure 5 also depicts the stiffness for each group. BS3-1 442.9 614.6 557.4 0.72 0.80
Whereas the maximum shear strength is achieved in speci- Group BS3-2 1421.8 1116.7 1132.6 1.27 1.26
mens with Grade 60 stirrups for Groups 1 and 2, stiffness in BS3 BS3-3 1178.5 1147.1 1199.8 1.03 0.98
such members is less than that in specimens with Grade 80
BS3-4 1490.6 1351.9 1663.1 1.10 0.90
stirrups. However, this is not the case for Groups 3 and 4. As
observed, the maximum shear strength is attained in speci- BS4-1 886.0 811.8 701.9 1.09 1.27
mens with Grade 80 stirrups, but stiffness in such members is Group BS4-2 1684.1 1439.2 1415.3 1.17 1.19
less than that in specimens with Grade 80 stirrups. Maximum BS4 BS4-3 1683.3 1477.2 1499.3 1.14 1.12
shear strength developed in the test is summarized in Table 2.
BS4-4 1795.8 1733.1 2078.2 1.04 0.86
A comparison of the maximum shear strength divided by the
cross section area is discussed in detail later. Notes: Vn-ACI is ACI 318-14 ; Vn-EC2 is EC2-04 ; 1 kN = 224.8 lb.
1 2

possibility of concrete web crushing prior to yielding of the


Shear failure mode
stirrup. Previous test data10-12 for beams less than 600 mm
The basic principle in primary design codes based on the
(23.6 in.) in cross section depth reported that beams with
truss model is that the yielding of a stirrup occurs prior to the
high-strength stirrups of 650 MPa (94.3 ksi) greater than
web crushing of concrete, leading to shear tension failure.
Grade 60 (420 MPa [60.0 ksi]) specified by ACI 318-141
For beams with high-strength stirrups, yield strain is linearly
failed in shear tension. The previous test results also eval-
increased as yield strength increases. This implies a strong
uated whether shear tension failure occurred in large-scale

ACI Structural Journal/January 2018 273


Fig. 6—Strain distribution of stirrups. (Note: 1 mm = 0.039 in.)
beams with high strength stirrups, having a large concrete
contribution to shear strength. To investigate the failure mode
of the current test specimens, stirrup strains were evaluated
and their distribution at five different loading levels is shown
in Fig. 6. The x- and y-axes in Fig. 6 represent the loca-
tion of stirrups between two supports and the stirrup strains,
respectively. Vmax in the legend refers to the maximum shear
strength obtained from the test. As observed, the stirrups
underwent yield strain in all specimens prior to reaching
the maximum shear strength. It is therefore assumed that all
specimens fail in shear tension required by ACI 318-14.1
For a further assessment of failure mode in relation
to cross section size, a comparison was made in terms
of using stirrup strains in both the current and previous Fig. 7—Comparison of normalized stirrup strains for
various levels of ρvfyt√fc′.

274 ACI Structural Journal/January 2018


Fig. 9—Normalized shear strength for various levels of
ρvfyt√fc′. (Note: 1 mm = 0.0394 in.; 1 MPa = 145 psi.)
of the previous test. It is thus observed from Fig. 7 that no
significant change occurs in shear failure mode, even for
large-scale members with high-strength stirrups.

DISCUSSION
Shear strength
Figure 8(a) shows the shear strength variation according
to cross section depth. For comparison, the normalized shear
strength ratio is used and summarized in Table 2. The normal-
ized shear strength ratio in the y-axis is the measured shear
strength Vn-test divided by the theoretical shear strength Vn-ACI
calculated using the ACI 318-141 equations: Eq. (22.5.1.1),
(22.5.6.1(a)), and (22.5.10.5.3). The strength reduction factor
of 1.0 and the actual measured yield strength of stirrups are
Fig. 8—Normalized shear strength for various section employed in the calculation of the theoretical shear strength.
depths. (Note: 1 mm = 0.0394 in.) For specimens without stirrups (BS1-1, BS2-1, BS3-1, and
test results.10-12 Figure 7 shows the stirrup strains for the BS4-1) shown in Fig. 8(a), while the normalized ratios in
current test with large-scale cross section and the previous Specimens BS1-1 and BS4-1 are greater than 1.0, those in
test10-12 with a section depth less than or equal to 600 mm Specimens BS2-1 and BS3-1 are not greater than 1.0. In
(23.6 in.). The y-axis in Fig. 7 represents the stirrup strain particular, although the normalized ratio of BS4-1 is greater
εpeak measured at the maximum shear strength divided by than 1.0, the margin is as low as 1.1 approximately. This
the yield strain of stirrup εy—that is, εpeak/εp ≥ 1 indicates implies that Eq. (22.5.1.1) of ACI 318-141 is likely to give an
yielding of stirrup. Shear failure mode is closely associated unconservative shear strength prediction for specimens with
with the yield strength of stirrup fyt, stirrup ratio ρv, and a cross section depth of not less than 850 mm (33.5 in.). In
compressive strength of concrete fc′. In the case where the general, the normalized ratio for specimens without stirrups
value of ρv fyt/√fc′ for a member is large, there is a strong tends to decrease as the cross section size increases.
possibility that the member fails in shear compression— The normalized ratios for all specimens with stirrups are
that is, concrete web crushing occurs prior to yielding of greater than 1.0. However, they asymptotically decrease as
stirrup. However, for the case of a small value of ρv fyt/√fc′, the cross section size increases. The normalized ratio was
the member is highly expected to fail in shear tension—that a particularly low value of 1.03 and 1.04 for Specimens
is, concrete web crushing occurs after yielding of stirrup. BS3-3 and BS4-4, respectively. The decreasing tendency in
In Fig. 7, the stirrup yield strength of fyt varies between 459 the normalized ratio reveals the unfavorable aspect whereby
and 750 MPa (66.6 and 100.1 ksi), and the maximum stirrup theoretical strength by ACI 318-141 is likely to exceed the
ratio satisfies the ACI 318-141 provision of ρv fyt/√fc′ ≤ 2/3. measured strength for specimens with a cross section depth
In Fig. 7, it is observed that εpeak/εy decreases as ρv fyt/√fc′ is greater than 1350 mm (53.1 in.). This suggests that the shear
increased. It is worth noting that, as shown in Fig. 7, shear equation by ACI 318-141 may lead to an erroneous overesti-
compression failure occurs in two specimens with both high- mation in the theoretical shear strength as the cross section
strength stirrup and high value of ρv fyt/√fc′. It is also note- size increases. This is due to the fact that the calculation
worthy that, as shown in Fig. 7, εpeak/εy is more influenced of shear strength provided by concrete Vc in the equation
by ρv fyt/√fc′ than by the cross section size. The variance of does not take into account the influence of size effect. The
εpeak/εy measured in large-scale specimens of the current test normalized ratio for specimens with Grade 60 stirrups is
exhibits a similar trend to that in the small-scale specimens generally greater than that for specimens with Grade 80 stir-

ACI Structural Journal/January 2018 275


Fig. 10—Shear-critical areas and schemes for diagonal crack measurement. (Note: Dimensions are in mm; 1 mm = 0.0394 in.)
rups. The margin of both ratios decreases as the cross section 78 specimens are employed in the comparison. Parameter
size increases. ranges vary widely: between 334 and 750 MPa (48.4 and
Further comparison is made in terms of using the predic- 108.8 ksi) for yield strength of stirrup; between 25.0 and
tions by EC2-04.2 Theoretical shear strength Vn-EC2 calcu- 81.4 MPa (3.6 and 11.8 ksi) for compressive strength of
lated by EC2-042 shear design equations and the normal- concrete; and between 350 and 1350 mm (13.8 and 53.1 in.)
ized shear strength ratio Vn-test/Vn-EC2 are also given in for cross section depth. Details of the specimens used in the
Table 2. In general, similar trends as in the case of compar- comparison are summarized in the Appendix.*
ison with ACI 318-141 are observed. However, the theoret- As observed in Fig. 9, the normalized shear strength ratio for
ical predictions by EC2-042 are generally greater than those specimens with a depth not greater than 600 mm (23.6 in.) is
by ACI 318-14.1 This is attributed to the fact that shear generally decreased as the value of ρv fyt/fc′ increases. More
design equations of EC2-042 account for shear resistance by specifically, while the ratio is 1.3 at 0.05 of ρv fyt/fc′, the ratio
stirrups only and the angle between the concrete compres- decreases to 1.05 at 0.1 of ρv fyt/fc′. In addition, the ratio for
sion strut and the beam axis perpendicular to the shear force. specimens with high-strength stirrups is lower than that
The limiting values of the angle are varied and can be found for specimens with normal-strength stirrups at an identical
in EC2-04.2 This leads to an increase of the shear predictions value of ρv fyt/fc′. Meanwhile, it is worth noting that the value
as the cross section depth increases and, thus, a decrease in of ρv fyt/fc′ is a relatively low level of 0.025 for the current test
the normalized shear strength ratio is prominent for beams specimens with Grade 80 stirrups and a cross section depth
with a large cross section depth. In particular, the normalized of not less than 850 mm (33.5 in.). Nonetheless, the normal-
shear strength ratio is less than 1 for specimens with a cross ized shear strength ratio of such specimens is approximately
section depth with 1350 mm (53.1 in.). This indicates that 1.1, which is less than that of specimens with a cross section
shear equation of EC2-042 is likely to overestimate the shear depth not greater than 600 mm (23.6 in.).
strength with a cross section depth greater than 1350 mm
(53.1 in.). Consequently, further assessment is required when Diagonal cracks
Grade 80 stirrups are employed in a large-scale member with The primary national design codes for reinforced concrete
a size greater than that in the present study. structures including ACI 318-141 limit the yield strength of
Figure 8(b) shows the measured shear strength provided by stirrups. One of the main reasons is to control crack width.
stirrups divided by the calculated shear strength provided by In particular, when high-strength stirrups are employed in a
stirrups using Eq. (22.5.10.5.3) of ACI 318-141 for cross section reinforced concrete structure, there is a high possibility that
depths. The measured shear strength provided by stirrups is diagonal crack width is increased. Shoya et al.26 revealed
obtained from the shear strength of a specimen with stirrups that the width can be increased as the section depth is
subtracted by that of such specimen without stirrups in each increased. In the present study, measurements for the width,
group. Figure 8(b) shows that the normalized ratio of shear angle of inclination, and the number of diagonal cracks are
strength provided by stirrups for members with Grade 80 stir- taken in shear-critical areas. The shear-critical areas are
rups is lower than that for such members with Grade 60 stirrups. assumed to be the portion of the concrete web between a line
Figure 9 shows a comparison of the relationship between equal to half the section depth from the loading point and
normalized shear strength ratio Vn-test/Vn-ACI and amount of a location extending the section depth from the line along
stirrups divided by the compressive strength of concrete
ρv fyt/fc′. Depicted in Fig. 9 are the previous test results8,9,25 *
The Appendix is available at www.concrete.org/publications in PDF format,
for beams with high-strength stirrups and a depth not greater appended to the online version of the published paper. It is also available in hard copy
than 600 mm (23 in.) and the current test results. A total of from ACI headquarters for a fee equal to the cost of reproduction plus handling at the
time of the request.

276 ACI Structural Journal/January 2018


Fig. 11—Measured diagonal crack width. (Note: 1 mm = 0.0394 in.)
the member axis. These areas are illustrated as the shaded shear critical areas has an angle of inclination between 20
areas in Fig. 10. When measuring the crack width, a straight and 60 degrees.
line perpendicular to a diagonal crack is drawn as shown in 2. In the case where the vertical length of a crack in the
Fig. 10. Then, Demec gauges and crack magnifiers are used shear critical areas is greater than or equal to d/6.
to measure the diagonal crack width at each loading level. Figure 11 shows the change of diagonal crack width
Cracks formed in the top and bottom cross sections are not measured in the shear-critical areas at various load levels.
included in the measurement because they are mostly flex- Diagonal crack barely occurred up to 40% of the load level.
ural cracks. In addition, cracks formed close to the supports Once the load level exceeded 40%, a diagonal crack formed
or loading points are also excluded. Thus, only diagonal in the concrete web, and the crack width increased as the
cracks occurring in the middle web of half the section depth load level increased. The crack width varied according to
are evaluated. Cracks satisfying the following conditions in the occurrence of the crack. In general, the earlier the occur-
the shear-critical areas are regarded as diagonal cracks: rence, the larger the crack width at the maximum load level.
1. In the case where a line between the starting and ending The difference between the maximum and minimum crack
points of cracks occurred in the upper and lower parts of the widths is approximately 1 mm (0.039 in.) in identical spec-

ACI Structural Journal/January 2018 277


Fig. 12—Average diagonal crack width at service load level.
(Note: 1 mm = 0.0394 in.)
imens. Figure 12 shows the average width of the diagonal
cracks measured at 60% of shear strength calculated by ACI
318-14.1 It is generally observed that the crack width and
spacing between the cracks increase as the stirrup spacing
increases from 130 to 180 mm (5.12 to 7.09 in.). In partic-
ular, the average diagonal crack width in specimens with
Grade 80 stirrups is larger than that in specimens with
Grade 60 stirrups. The width increases as the cross section
depth increases. For specimens with a cross section depth
not greater than 600 mm (23.6 in.) in the previous test,10-13
the width of such specimens with Grade 80 stirrups is not
greater than 0.41 mm (0.017 in.), which is the upper bound
recommended by ACI 224R-0115 at the service load level.
However, for a cross section depth not less than 1200 mm
(47.2 in.) in the current test, the width exceeds 0.41 mm
(0.017 in.). This implies that large-scale specimens with
high-strength stirrups may exceed 0.41 mm (0.017 in.) of
the crack width. Fig. 13—Average angle of inclination and average number
Figures 13(a) and 13(b) show the average angle of inclina- of diagonal cracks at service load level. (Note: 1 mm =
tion and the number of diagonal cracks measured in the shear 0.0394 in.)
areas, respectively. As observed, a number of diagonal cracks
increased as the cross section depth was increased. However, ment, additional test results are required respect to a number
the angle of inclination of diagonal cracks slightly reduced of the diagonal cracks and the angle of the cracks in relation
as the cross section depth was increased. Although the shear to cross section depth.
span-depth ratio is identical in all specimens, the area of
concrete web is relatively increased as cross section depth SUMMARY AND CONCLUSIONS
increases. Thus, numerous diagonal cracks passing through The applicability of Grade 80 high-strength stirrups to
the relatively large area of concrete web increased when the reinforced concrete members has been actively studied over
cracks formed due to force redistribution. The angle of incli- the last decade. However, most previous works focused
nation of diagonal cracks in group BS1 with the smallest on members with relatively small-scale cross sections,
cross section depth increased because the cracks initiated at and very few studies have been carried out for large-scale
a certain distance away from support and propagated toward members with Grade 80 stirrups. In this paper, evaluation
a loading point. However, the angle of cracks in Group BS4 was performed with respect to the shear behavior of large-
with the largest cross section depth decreased because the scale reinforced concrete beams with Grade 80 stirrups. For
cracks formed initially near of support and extended toward this purpose, a total of 16 beam specimens were tested in the
a loading point. In addition, the angles of cracks are closely current study. The conclusions derived from the current test
associated with a stirrup spacing together with the cross results are presented as follows:
section depth. The angle of cracks decreased as the stirrup 1. All of the large-scale reinforced concrete beams with
spacing increased. The above observation is derived from Grade 80 stirrups failed in shear tension—that is, yielding of
the present test results. Subsequently, for a further assess- stirrups occurs prior to concrete web crushing. The measured
shear strength is greater than the theoretical shear strength

278 ACI Structural Journal/January 2018


calculated by ACI 318- 141 in all specimens, even for speci- Ho-Young Kim is a PhD Candidate in the Department of Civil Engineering
at Yeungnam University, Gyongsan, South Korea, where he received his BE
mens of 1350 mm (53.1 in.) section depth in which Grade 80 and MS in civil engineering.
stirrups were employed. It is thus promising that Grade 80
high-strength stirrups can be employed in large-scale rein- ACKNOWLEDGMENTS
forced concrete beams. It is, however, worth noting that the This research was supported by the Korea Hydro & Nuclear Power Co.,
normalized shear strength ratio Vn-test/Vn-ACI decreases as the LTD (KHNP, contract number: L16S140000) and the Basic Science Research
Program through the National Research Foundation of Korea (NRF) funded
cross section depth increases. by the Ministry of Education (grant number: 2015R1D1A1A01056612).
2. Previous and current test results on 78 reinforced
concrete beams with high-strength stirrups indicate that the NOTATION
normalized shear strength ratio of large-scale beams is lower b = width of beam cross section
than that of relatively small-scale beams. In the present test d = effective depth of beam cross section
fc′ = compressive strength of concrete
however, a low level of 0.025 for ρv fyt/fc′ is employed in the fyl = yield strength of longitudinal tensile reinforcement
large-scale beams. Hence, additional investigation is needed fyt = yield strength of stirrup
to determine the applicability of high-strength stirrups to h = depth of beam cross section
s = stirrup spacing
large-scale beams with a relatively high level of ρv fyt/fc′. V = shear strength
3. The average diagonal crack width reduces as the Vc = shear strength provided by concrete
number of stirrups is increased. The width in beams with Vn-ACI = shear strength calculated by ACI 318-14
Vn-test = measured shear strength by test
Grade 80 stirrups is larger than that in beams with Grade 60 Vs = shear strength provided by stirrups
stirrups and increased as the stirrup spacing was increased. ρv = stirrup ratio
In addition, the width increased as the cross section depth is ρv fyt = amount of stirrups
increased. In particular, for beams with a cross section depth
of 1350 mm (53.1 in.), the width is greater than 0.41 mm REFERENCES
1. ACI Committee 318, “Building Code Requirements for Structural
(0.017 in.), which is the crack width recommended by ACI Concrete (ACI 318-14) and Commentary (ACI 318R-14),” American
224R-0115 at the service load level. Concrete Institute, Farmington Hills, MI, 2014, 519 pp.
4. The angle of inclination of diagonal cracks is reduced 2. Comite European de Normalisation (CEN), “Eurocode 2: Design of
Concrete Structures, Part 1-1 General Rules and Rules for Buildings (BS
as the cross section depth increases and the stirrup ratio is EN 1992-1-1),” Lausanne, Switzerland, 2004, 211 pp.
reduced. The number of diagonal cracks increased as the 3. CSA Committee A23.3-04, “Design of Concrete Structures for Build-
cross section depth is increased. Although it is rather prema- ings (CAV3-A23.3-04),” Canadian Standards Association, Toronto, ON,
Canada, 2004, 232 pp.
ture to generalize a tendency, the aforementioned phenomena 4. Hosny, A.; Seliem, H. M.; Rizkalla, S.; and Zia, P., “Development
can be used as useful information regarding the evaluation Length of Unconfined Conventional and High-Strength Steel Reinforcing
of shear behavior on large-scale beams, in the lack of avail- Bars,” ACI Structural Journal, V. 109, No. 5, Sept.-Oct. 2012, pp. 655-664.
5. Briggs, M.; Miller, S.; Darwin, D.; and Browning, J., “Bond Behavior
able test data on the angle of inclination and the number of of Grade 100 AST A1035 Reinforcing Steel in Beam-Splice Specimens,”
diagonal cracks for such members. SL Report 07-1, The University of Kansas Center for Research, Lawrence,
Conclusions discussed previously are derived from a total KS, 2007, 83 pp.
6. Aoyama, H., Design of Modern Highrise Reinforced Concrete Struc-
of 16 specimens. Subsequently, additional tests are required tures, Imperial College Press, London, UK, 2001, 442 pp.
further to generalize a practicability of the Grade 80 stirrups 7. Li, Y.-A., and Hwang, S.-J., “Prediction of Shear Strength for Rein-
and propose some suggestion in design formulas. In partic- forced Concrete Short Columns Using High Strength Material,” The
15th Korea-Japan-Taiwan Joint Seminar on Earthquake Engineering for
ular, influence of concrete compressive strength, stirrup Building Structures, Taipei, Taiwan, 2013, pp. 41-50.
spacing, and ratio on shear behavior of large-scale beams 8. Ou, Y. C., and Kurniawan, D. P., “Effect of Axial Compression on
needs to be evaluated in the further tests. Shear Strength of Reinforced Concrete Columns with High-Strength Steel
and Concrete,” The 15th Korea-Japan-Taiwan Joint Seminar on Earthquake
Engineering for Building Structures, Taipei, Taiwan, 2013, pp. 105-114.
AUTHOR BIOS 9. Choi, W.-S.; Park, H.-G.; Chung, L.; and Kim, J. K., “Experimental
ACI member Jung-Yoon Lee is a Professor in the School of Civil, Architec- Study for Class B Lap Splice of 600 MPa (87 ksi) Reinforcing Bars,” ACI
tural Engineering, and Landscape Architecture at Sungkyunkwan Univer- Structural Journal, V. 111, No. 1, Jan.-Feb. 2014, pp. 49-58.
sity, Seoul, South Korea. His research interests include shear behavior and 10. Lee, J.-Y., and Choi, I.-J., “Shear Behavior of Reinforced Concrete
seismic design of reinforced and prestressed concrete structures. Beams with High-Strength Stirrups,” ACI Structural Journal, V. 108,
No. 5, Sept.-Oct. 2011, pp. 620-629.
ACI member Jae-Hoon Lee is a Professor in the Department of Civil Engi- 11. Lee, J.-Y.; Lee, D. H.; Lee, J.-E.; and Choi, S.-H., “Shear Behavior
neering at Yeungnam University, Gyongsan, South Korea. He received and Diagonal Crack Width for Reinforced Concrete Beams with High
his BE and MS from Seoul National University, Seoul, South Korea, and Strength Shear Reinforcement,” ACI Structural Journal, V. 112, No. 3,
his PhD from the University of Wisconsin–Madison, Madison, WI. His May-June 2015, pp. 323-333. doi: 10.14359/51687422
research interests include behavior, analysis, and design of reinforced and 12. Lee, J.-Y.; Choi, S.-H.; and Lee, D. H., “Structural Behavior of Rein-
prestressed concrete structures. forced Concrete Beams with High Yield Strength of Stirrups,” Magazine
of Concrete Research, V. 68, No. 23, 2016, pp. 1187-1199. doi: 10.1680/
Do Hyung Lee is a Professor in the Department of Civil, Environmental jmacr.15.00344
and Railroad Engineering at Pai Chai University, Daejeon, South Korea. 13. Munikrishma, A.; Hosny, A.; Rizkalla, S.; and Zia, P., “Behavior of
He received his BSC and MSc from Hanyang University, South Korea, and Concrete Beams Reinforced with ASTM A1035 Grade 100 Stirrups under
his MSc and PhD from Imperial College London, London, UK. His research Shear,” ACI Structural Journal, V. 108, No. 4, July-Aug. 2011, pp. 34-41.
interests include seismic design and performance of concrete structures: 14. Seliem, H.; Hosny, A.; Rizkalla, S.; Zia, P.; Briggs, M.; Miller, S.;
analytical and experimental studies. Darwin, D.; Browning, J.; Glass, G.; Hoyt, K.; Donnelly, K.; and Jirsa, J.,
“Bond Characteristics of ASTM A1035 Steel Reinforcing Bars,” ACI Struc-
Seong-Jun Hong is an Engineer at Samsung C&T, Engineering & tural Journal, V. 106, No. 4, July-Aug. 2009, pp. 530-539.
Construction Group, South Korea. He received his MS from Sungkyunkwan 15. ACI Committee 224, “Control of Cracking in Concrete Structures
University. His research interests include the structural analysis and design (ACI 224R-01),” American Concrete Institute, Farmington Hills, MI, 2001,
of concrete structures. 46 pp.

ACI Structural Journal/January 2018 279


16. Kani, G. N. J., “How Safe Are our Large Reinforced Concrete 22. Adebar, P., “Diagonal Cracking and Diagonal Crack Control in Struc-
Beams?” ACI Journal Proceedings, V. 64, No. 3, Mar. 1967, pp. 128-141. tural Concrete” Design and Construction Practices to Mitigate Cracking,
17. Zararis, P., and Papadakis, G. C., “Diagonal Shear Failure and Size SP-204, E. G. Nawy, F. Barth, and R. Frosch, eds., American Concrete Insti-
Effect in RC Beams without Web Reinforcement,” Journal of Structural tute, Farmington Hills, MI, 2001, pp. 85-116.
Engineering, ASCE, V. 127, No. 7, 2001, pp. 733-742. doi: 10.1061/ 23. Brooms, B., “Crack Width and Crack Spacing in Reinforced
(ASCE)0733-9445(2001)127:7(733) Concrete Members,” ACI Journal Proceedings, V. 62, No. 10, Oct. 1965,
18. Bentz, E. C., “Shear Strength of Members without Transverse Rein- pp. 1237-1256.
forcement,” Structures Congress, 2009, pp. 1604-1611. 24. Hassan, H. M.; Farghaly, S. A.; and Ueda, T., “Displacements at
19. Bažant, Z. P., and Yu, Q., “Designing Against Size Effect on Shear Shear Crack in Beams with Shear Reinforcement under Static and Fatigue
Strength of Reinforced Concrete Beams without Stirrups: I. Formulation,” Loadings,” Proceedings of the JSCE, V. 19, Aug. 1991, pp. 215-222.
Journal of Structural Engineering, ASCE, V. 131, No. 12, 2005, pp. 1877- 25. Zakaria, M.; Ueda, T.; Wu, Z.; and Meng, L., “Experimental Inves-
1885. doi: 10.1061/(ASCE)0733-9445(2005)131:12(1877) tigation on Shear Cracking Behavior in Reinforced Concrete Beams with
20. Yu, Q., and Bažant, Z. P., “Can Stirrups Suppress Size Effect on Shear Shear Reinforcement,” Journal of Advanced Concrete Technology, V. 7,
Strength of RC Beams?” Journal of Structural Engineering, ASCE, V. 137, No. 1, 2009, pp. 79-96. doi: 10.3151/jact.7.79
No. 5, 2011, pp. 607-617. doi: 10.1061/(ASCE)ST.1943-541X.0000295 26. Shoya, T.; Iguro, M.; Nojiri, Y.; Akiyama, H.; and Okada, T., “Shear
21. Regan, P. L., and Placas, A., “Shear Failure of Reinforced Concrete Strength of Large Reinforced Concrete Beams,” Fracture Mechanics:
Beams,” ACI Journal Proceedings, V. 68, No. 10, Oct. 1971, pp. 763-773. Application to Concrete, SP-118, V. C. Li and Z. P. Bažant, eds., American
Concrete Institute, Farmington Hills, MI, pp. 259-280.

280 ACI Structural Journal/January 2018


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 115-S23

Use of Anchored Carbon Fiber-Reinforced Polymer Strips


for Shear Strengthening of Large Girders
by William A. Shekarchi, Wassim M. Ghannoum, and James O. Jirsa

The objective of this research was to investigate the feasibility of


increasing the shear capacity of large reinforced concrete girders
using anchored carbon fiber-reinforced (CFRP) strips. In an exper-
imental study consisting of nine tests, the effects of unidirectional
and bidirectional CFRP layouts, anchored and fully wrapped
systems, retrofitting uncracked and precracked sections, and
loading conditions on the shear performance of large girders with
relatively wide webs were investigated.
Test results indicated that CFRP anchors were able to develop
the fracture strength of the CFRP strips despite some anchors being
placed in flexural tension zones. The unidirectional CFRP layouts
were able to improve the shear capacity by as much as 56% while
the bidirectional layouts improved the inclined cracking shear load
by as much as 22%. Moreover, the measured lower-bound average
CFRP fracture strain was found to be significantly larger (0.007) Fig. 1—Shear crack in bent cap girder. (Courtesy of the
than the maximum permitted effective strain (0.004) currently Texas Department of Transportation.)
recommended for design.
to buckling when a continuous anchorage plate is placed
Keywords: carbon-fiber reinforced polymer (CFRP); CFRP anchors; rein- within a member’s compression region and slippage of
forced concrete (RC); shear behavior; shear strengthening. the CFRP strips due to inadequate clamping forces. More
recently, research on the use of CFRP spike anchors, hence-
INTRODUCTION forth referred to as CFRP anchors, has begun to grow due
Many structures in the United States are reaching their to the ability of properly detailed anchors to develop the
expected service life due to increasing load demands, occu- full tensile capacity of the CFRP strips and to their ease
pancy requirements, exposure to corrosive environments, of installation.8-14
and sometimes-inadequate structural designs. The costs Many of the experimental studies related to CFRP shear
associated with replacing a large element in a bridge or a strengthening have been conducted on small-scale, thin-
building are often too high to be a feasible option. Effec- webbed beams that do not meet minimum transverse steel
tive and efficient rehabilitation techniques offer an attractive reinforcement requirements.2,3,11 Hence, the observed shear
solution for rehabilitating aging or inadequate elements if strength gains of these members likely do not represent the
the cost of the installation and disruption to the use of the shear strength gains of large in-place members. Therefore,
structure can be minimized. the feasibility of improving the shear capacity of large-scale
Carbon fiber-reinforced polymer (CFRP) is an ideal mate- relatively wide-webbed reinforced concrete (RC) bent cap
rial for rehabilitating existing members due to its high tensile girders using CFRP strips and CFRP anchors was studied
strength-to-weight ratio, adaptability to a variety of member in this project. Shear distress in a bridge bent cap girder is
shapes, and rapid installation that minimizes disruption to the shown in Fig. 1. The effects of the following variables on the
use of the structure. However, without proper anchorage, the shear performance of the square 32 in. (813 mm) deep and
interfacial bond between the CFRP and the surface concrete wide bent cap girders were studied: 1) unidirectional and
is lost at CFRP stresses that are less than half of the CFRP bidirectional CFRP layouts; 2) anchored and fully wrapped
fracture stress.1-4 Consequently, debonding failures limit the systems; 3) retrofitting uncracked and precracked sections;
efficiency and shear contribution of the CFRP material. and 4) loading conditions. The effects of additional variables
Rather than rely solely on the interfacial bond strength, have been studied in other parts of the experimental program
anchorage for CFRP strips has been studied for cases where and are discussed in Kim et al.10 and Jirsa et al.14
the CFRP strips cannot be fully wrapped due to obstruc-
tions (that is, a bent cap girder supporting I-beams or beams
that are cast monolithically with floor slabs). Much of the ACI Structural Journal, V. 115, No. 1, January 2018.
research conducted on anchorage systems was based on MS No. S-2017-093, doi: 10.14359/51701092, was received March 14, 2017, and
reviewed under Institute publication policies. Copyright © 2018, American Concrete
the use of mechanical anchors that clamp the CFRP strips Institute. All rights reserved, including the making of copies unless permission is
obtained from the copyright proprietors. Pertinent discussion including author’s
in place.5-7 Clamping anchorage systems are susceptible closure, if any, will be published ten months from this journal’s date if the discussion
is received within four months of the paper’s print publication.

ACI Structural Journal/January 2018 281


Fig. 2—Cross sections of bent cap girder. (Note: 1 in. = 25.4 mm.)

Fig. 3—Description of tests and nomenclature.


RESEARCH SIGNIFICANCE curvature specimen had a longitudinal reinforcement ratio
Reinforced concrete members may experience shear defi- (ρ) of 1.9% for an effective depth (d) of 28 in. (711 mm).
ciencies due to increasing load demands, inadequate struc- As a result, the single and double curvature specimens had
tural designs, or inadvertent damage. In most cases, it is similar strains in the longitudinal reinforcement at the same
more economical to strengthen rather than replace members. applied shear load. The large longitudinal reinforcement
Externally bonded CFRP strips present an ideal approach ratios were necessary to ensure that the shear capacity of the
for strengthening existing members due to their high tensile strengthened girders did not exceed the applied shear load
strength-to-weight ratio and expedited installation times. that would cause flexural yielding.
However, the efficiency of the CFRP layout depends on the Regardless of the test configuration, the spacing of the
anchorage system. In this study, CFRP anchors were used to transverse steel reinforcement in the test span was deter-
develop the strength of the U-wrap (or discontinuous) CFRP mined such that the specimens met the minimum trans-
strips. Substantial shear strength gains were observed for the verse reinforcement requirements in accordance with
large-scale bent cap girders using anchored CFRP strips. ACI and AASHTO specifications.15,16 All the specimens
were designed to have span-depth ratios (av/d) equal to
TEST CONFIGURATION 3.0, where the shear span (av) was measured as the clear
Specimen design distance between the loading plate and reaction bearing
To meet the constraints of the test facilities, the test spec- pad. However, it should be noted that the span-depth ratio
imens had a 32 in. (813 mm) deep by 32 in. (813 mm) wide for a concrete member loaded in double curvature is gener-
cross section and were 27 ft 8 in. (8433 mm) long. ally implied to be reduced due to the existence of a point of
The loading and support conditions of in-place bent inflection between the positive and negative moments.15
cap girders generally result in double, or reverse, curva- Typical cross sections are shown in Fig. 2. Note that
ture. Hence, the members will have positive and negative the outer bundled bars in the double curvature specimen
moments along the shear span. Seven of the test specimens (Fig. 2(b)) were moved inward for the last three tests in
were subjected to loading that produced double curvature, an attempt to mitigate bond-shear cracks caused by a high
and two tests were conducted using loading that resulted in strain gradient in the compression reinforcement, as will be
single curvature. Thus, the experimental results from the discussed later.
double curvature tests could eventually be compared with
results from other experimental studies.10,14 Nomenclature
The proportion of the longitudinal reinforcement varied A nomenclature was developed to help define the param-
between the single and double curvature specimens due eters in each test. A graphical representation of the nomen-
to differences in the magnitude of the maximum applied clature can be seen in Fig. 3. The first term indicates how the
moments. Specifically, the maximum moment in the double girder was loaded. The second term indicates if the girder
curvature specimen was approximately half of the maximum was previously tested and, thus, precracked. The third and
single curvature moment for the same applied shear force fourth group of terms indicate whether vertical or horizontal
in the test span. Consequently, the single curvature spec- strips were used and if so, how the CFRP strips were devel-
imen had a longitudinal reinforcement ratio (ρ) of 3.6% for oped. The nomenclature and a short description of each test
an effective depth (d) of 27 in. (686 mm) while the double is provided in Table 1.

282 ACI Structural Journal/January 2018


Fig. 4—CFRP layouts. (Note: 1 in. = 25.4 mm.)

Table 1—Specimen description CFRP anchors


A significant amount of research complementary to this
Nomenclature Description
project has been conducted on the optimization of CFRP
S-U-VN-HN Control anchors.8-10,14,17,18 In particular, Jirsa et al.14 proposed
S-C-VA-HN Anchored unidirectional, cracked comprehensive CFRP anchor design recommendations that
D-U-VN-HN Control determine the required amount of CFRP anchor material,
anchor hole diameters, chamfer bend radii, anchor fan bond
D-C-VF-HN Fully wrapped unidirectional, cracked
lengths, and anchor embedment depths for user fabricated
D-U-VF-HN Fully wrapped unidirectional, uncracked and premanufactured CFRP anchors.
Fully wrapped vertically and anchored horizontally, For all tests, 10 in. (254 mm) long premanufactured CFRP
D-U-VF-HA
uncracked anchors were used. The CFRP anchors on the boundary of
D-U-VN-HN*† Redundant control the vertical and horizontal CFRP strips provided an anchor-
D-C-VA-HN †
Anchored unidirectional, cracked
to-strip cross-sectional area ratio of 2.9 while the interme-
diate horizontal strip anchors provided a ratio of 2.4 after
D-U-VA-HA† Anchored bidirectional, uncracked
the intermediate anchors had been divided into two fans
*
Redundant double curvature control test. that were splayed in opposing directions over the strips,

Modified cross section and test setup. as shown in Fig. 4(c) and 4(d). Generally, providing an
anchor-to-strip cross-sectional area ratio of 2.0 was found
CFRP layouts to develop the tensile strength of the CFRP strips.10,14 Note
Four CFRP layouts were used to strengthen the bent cap that the intermediate horizontal anchors were spaced at a
girders in shear, as shown in Fig. 4. The CFRP layouts were distance approximately equal to the effective depth of the
designed such that each layout had the same cross-sectional section to minimize the elongation of the horizontal CFRP
area of CFRP crossing an assumed 45-degree shear crack, strips, which in turn was likely to reduce shear crack widths.
that is, half of the vertical CFRP strips in the unidirectional To simplify the anchor installation procedure used by Kim et
layouts were removed and placed in the horizontal direction al.,12 the vertical and horizontal CFRP strips were discontinued
for the bidirectional layouts. just short of the 9/16 in. (14 mm) diameter and 4 in. (102 mm)
All the CFRP strips in this experimental study had a deep boundary anchor holes, as shown in Fig. 5(a). Conversely,
nominal 3 in. (76 mm) width. Each CFRP strip location was the horizontal strips were placed over the 3/4 in. (19 mm) diam-
composed of two individual CFRP strips spliced together at eter and 4 in. (102 mm) deep intermediate anchor holes similar
the top and/or bottom of the girder. For instance, the anchored to Kim et al.,12 making insertion of the anchors through the
U-wrap (or discontinuous) CFRP strips were spliced only on strip troublesome. After installing the CFRP anchors, two
the bottom of the girders whereas the fully wrapped CFRP patches were laid over the holes and anchors in a perpendic-
strips were spliced on the top and bottom of the girders. As ular and parallel fiber direction to mitigate stresses within
a result, it was easier to ensure that air entrapped under the the anchors. The boundary and intermediate CFRP anchor
strip was removed. All the CFRP strip splices had at least an dimensions are shown in Fig. 5.
8 in. (203 mm) overlap length, which provided a bond stress By using the double curvature loading conditions, repre-
less than the CFRP manufacturer’s maximum allowable sentative of many in-place bent cap girder conditions, the
bond stress. Moreover, all sharp edges (corners and anchor performance of the CFRP anchors in a tension region could
holes) were rounded to a bend radius of 0.5 in. (13 mm) to be investigated because approximately half of the CFRP
mitigate stress concentrations in the CFRP. anchors were placed within a flexural tension zone. Many
of the CFRP anchors were installed through shear or bond-
shear cracks in the precracked specimens.

ACI Structural Journal/January 2018 283


Table 2—CFRP laminate properties reported by
manufacturer
Typical test value Design value
Tensile elastic modulus 13,900 ksi 11,900 ksi
Tensile ultimate strength 143 ksi 121 ksi
Ultimate elongation at fracture 1% 0.85%
Thickness 0.02 in. 0.02 in.

Notes: 1 ksi = 6.9 MPa; 1 in. = 25.4 mm.

(in psi units) and the bidirectional specimens had a vertical


contribution of 83 psi (0.57 MPa) or 1.2√fc′ (in psi units). As a
point of reference, ACI 318-1415 limits the shear contribution
of the transverse reinforcement to 8√fc′ (in psi units).

Test setup
In the single curvature test setup shown in Fig. 6(a), the
applied shear in the test span was measured directly at the
south support using two load cells. In the double curvature
test setup shown in Fig. 6(b), an additional loading point
produced positive and negative moments in the test span.
Approximately equal maximum moments at both ends of the
Fig. 5—CFRP anchor detail. (Note: 1 in. = 25.4 mm.) test span were achieved by using a load maintainer, which
produced a load at the north ram approximately twice that at
Material properties the south ram. The shear in the test span was calculated as
A ready mixed concrete supplier provided a concrete the south reaction minus the load applied by the south ram.
mixture with a test day strength of approximately 4500 psi Larger loading plates and reaction bearing pads were
(31 MPa). The longitudinal reinforcement was ASTM A615 used in the double curvature test setup compared to the
Grade 75 No. 11 reinforcing bars whereas the transverse single curvature test setup to prevent concrete crushing of
reinforcement was ASTM A615 Grade 60 No. 5 reinforcing the direct struts and nodes in the transfer spans outside of
bars. The average measured yield strength of the transverse the test span, which was between the loading and reaction
reinforcement was 61 ksi (421 MPa). The yield strength of points. The large loading plates and reaction bearing pads
the longitudinal reinforcement was not measured because are representative of typical dimensions of prestressed
the reinforcement did not yield. The CFRP laminate used I-beams that rest on top of bent cap girders and columns that
in this study had unidirectional fibers and exhibited a linear- support the girders.
elastic behavior until fracture. The CFRP laminate proper- The number of load cells used in the south reaction of the
ties (parallel to the fiber direction), as reported by the CFRP double curvature test setup was changed from four to two so
manufacturer, are shown in Table 2. that the rotation of the member could be better accommodated.
At the same time, the cross-sectional detail of the double
Transverse reinforcement comparison curvature specimens was altered as shown in Fig. 2(b). The
Traditionally, the total amount of transverse reinforce- changes to the cross section of the specimens and the south
ment is reported as a reinforcement ratio. However, a trans- reaction did not affect the shear capacity of the specimens.
verse reinforcement ratio is meaningless because the steel Load was applied to the control and strengthened speci-
and CFRP reinforcement have different material properties mens in stages until a loss of shear strength (that is, shear
and stress-strain relationships. Instead, the vertical shear failure) was experienced or the load-deflection curve indi-
stress contribution can be used as a surrogate for the trans- cated an increase in deflection with no change in the load.
verse reinforcement ratio because the material stress is Upon reaching the peak shear load, the control specimens
taken into account. were immediately unloaded. The control specimens were
Because CFRP anchors were provided, the U-wrap CFRP then strengthened. The behavior and performance of the bent
layouts were assumed to be equivalent to fully wrapped cap girders were monitored using strain gauges on the longi-
CFRP layouts. Consequently, the effective stress of both tudinal and transverse reinforcement, linear potentiometers
CFRP layouts was assumed, at the specimen design stage, at the locations of the rams and reactions, and an optical
to be 40% of the CFRP fracture stress as proposed by ACI measurement system to measure three-dimensional (3-D)
440.2R-081 for fully wrapped layouts. displacements and element strains on the surface of the
Using this approach, the control specimens had a vertical concrete and CFRP strips. The determination of the surface
shear stress contribution of 66 psi (0.46 MPa) or 1.0√fc′ (in psi displacements and element surface strains is discussed
units), while the unidirectional specimens had a contribution in-depth by Sokoli et al.19 and Shekarchi.20
from both steel and CFRP of 100 psi (0.69 MPa) or 1.5√fc′

284 ACI Structural Journal/January 2018


Fig. 6—Test configuration. (Note: 1 ft = 304.8 mm; 1 in. = 25.4 mm.)

Table 3—Summary of test results


Average concrete Cracking shear, Cracking shear Peak shear
Test name strength, ksi kip Peak shear, kip increase, % increase, % Crack angle, deg Failure mode*
S-U-VN-HN 4.71 165 295 — — 25 DT
S-C-VA-HN 4.71 — 365 — 24% 25 SF
D-U-VN-HN 4.40 148 293 — — 34 DT
D-C-VF-HN 4.40 — 424 — 45% 34 SF
D-U-VF-HN 4.41 161 456 9% 56% 31 SF
D-U-VF-HA 5.11 180 446 22% 52% 40 SF
D-U-VN-HN †
4.56 149 313 — — 31 DT
D-C-VA-HN 4.56 — 427 — 37% 31 SF
D-U-VA-HA 4.58 177 367 19% 17% 26 SF
*
Failure mode: DT is diagonal tension, and SF is CFRP strip fracture.

Redundant double curvature control test.
Notes: 1 ksi = 6.9 MPa; 1 kip = 4.45 kN, 1 deg = 0.017 rad.

EXPERIMENTAL RESULTS AND DISCUSSION In the subsequent discussion, an overview of the shear
As predicted, all the specimens failed in shear. In all the behavior of a typical uncracked member strengthened using
strengthened specimens, the CFRP strips that crossed the an anchored CFRP layout is presented. The behavior of an
critical shear crack fractured when the peak shear capacity uncracked and cracked member is discussed. Finally, the
was reached. Such failures represent the optimal failure effects of unidirectional and bidirectional CFRP layouts,
mode because the full strength of the CFRP strips was used. anchored and fully wrapped strips, strengthening of
Fracture of the CFRP strips demonstrated the viability of uncracked and cracked sections, and loading conditions on
using CFRP anchors. Shear strength gains up to 56% were the shear performance of the bent cap girders are discussed.
observed for the strengthened specimens relative to compa-
rable control specimens, as shown in Table 3.

ACI Structural Journal/January 2018 285


Fig. 7—Single curvature applied shear-displacement
response. (Note: 1 kip = 4.45 kN; 1 in. = 25.4 mm.)
Overview of shear behavior of strengthened
specimens
The following overview is based on observations herein
and available literature.10,20-24 Prior to the formation of a
shear crack, the concrete resists the applied shear force.
Immediately after a shear crack forms, a fraction of the
applied shear force is distributed to the transverse steel and
CFRP reinforcement. The force in the CFRP reinforcement
causes the strips to debond locally across the shear crack. As
the member is loaded further, the shear crack grows and the
CFRP strains increase causing the debonding of the strips
to spread outward from the shear crack toward the points of
anchorage. As the shear crack widens, the beneficial effects
of aggregate interlock are reduced. Once the transverse steel
reinforcement crossing the shear crack yields, any additional
applied shear force is resisted by the concrete and CFRP,
causing the rate of CFRP debonding to accelerate until the
Fig. 8—Double curvature applied shear-displacement
strips have fully debonded between the points of anchorage.
response. (Note: 1 kip = 4.45 kN; 1 in. = 25.4 mm.)
Hence, the anchors need to resist and transfer the strip forces
into the concrete. Adequately designed anchorage systems control specimens had lower cracking shear loads than the
enable the CFRP strips to fracture at the peak shear capacity single curvature control specimen even though the speci-
of the member, which is immediately followed by the loss of mens were designed similarly (Table 3). The cracking load
aggregate interlock. of the bidirectionally strengthened double curvature speci-
The only difference in behavior between an uncracked and mens was improved by as much as 22%.
cracked member is that the CFRP strains, and thus debonding, The addition of the transverse CFRP reinforcement
begin to increase immediately as the shear force is applied. improved the shear stiffness of the bent cap girder speci-
The degradation of bond between the concrete and CFRP mens after the formation of the shear crack. For instance,
strips in the uncracked and cracked members implies that the strengthened precracked specimen (S-C-VA-HN) had a
a strain compatibility between the two materials does not larger initial stiffness compared to the post-cracking shear
exist. However, displacement compatibility may be main- stiffness of the control specimen (S-U-VN-HN). A similar
tained after the anchored CFRP strips have fully debonded. trend was expected in the double curvature specimens but
bond-shear cracks formed along the compression flex-
Applied shear-displacement response ural reinforcement and reduced the stiffness of the control
The applied shear-displacement response of the bent cap specimens near the peak shear capacity. Note in Fig. 8 that
girder specimens offers insight into cracking loads, relative the strengthened double curvature specimens exhibited a
shear stiffness, and ultimate capacities. The relationship substantial change in stiffness when the transverse steel rein-
between the applied shear and displacement under the point forcement yielded.
load in the test span for the single and double curvature The strengthened precracked single curvature specimen
specimens are shown in Fig. 7 and 8, and the test results are (S-C-VA-HN) exhibited a 24% increase in the peak shear
summarized in Table 3. capacity, whereas the strengthened double curvature spec-
In this study, the inclined cracking shear load was defined imens exhibited up to a 56% increase in the peak shear
as the shear load that caused the shear-carrying mechanism capacity. The disparity in strength gains between the single
to change (that is, the steel transverse reinforcement began and double curvature specimens was likely the result of
to contribute to the shear resistance). The double curvature some damage due to loading the control single curvature

286 ACI Structural Journal/January 2018


Fig. 9—Double curvature midheight shear crack width
comparison. (Note: 1 kip = 4.45 kN; 1 in. = 25.4 mm.)
specimen (S-U-VN-HN) beyond the peak shear capacity (as
shown in Fig. 7), which was then strengthened and tested as
S-C-VA-HN.

Unidirectional and bidirectional CFRP layouts


The unidirectional layout provided a marginally larger
strength gain compared to the bidirectional layout even
though both layouts used a similar amount of CFRP on
the surface of the web. The better material use efficiency
of the unidirectional layout compared to the bidirectional
layout was likely due to the shallow shear crack angles
at the ultimate shear capacity, which were consistently
less than 45 degrees (Table 3). Consequently, the vertical
CFRP strips were able to develop fracture strength before
the horizontal strips. However, it should be mentioned that
the shear capacity of the anchored bidirectional specimen
Fig. 10—Uniform strain distribution. (Note: 1 in. = 25.4 mm.)
(D-U-VA-HA) was lower than expected due to observed
undesirable strain concentrations between overlapping and in the vertical and horizontal directions with the bidirec-
adjacent CFRP anchors, as shown in Fig. 4(d), which frac- tional layout. The increased shear crack restraint may have
tured the horizontal strips prior to the vertical strips. improved the concrete shear contribution.
Using the optical measurement system, the width of the
principal shear crack (wc) was estimated to determine if the Anchored and fully wrapped systems
crack width influenced the performance of the unidirec- Specimen D-C-VA-HN (with anchored vertical strips)
tional and bidirectional layouts. The width of the principal and D-C-VF-HN (with fully wrapped vertical strips) exhib-
shear crack at midheight of the section was used as a means ited similar shear strength capacities even though some of
of consistently comparing crack widths between tests. the CFRP anchors were placed through preexisting cracks.
However, the midheight and maximum shear crack widths The relative efficiency of the anchored and fully wrapped
had similar trends. systems can be determined by investigating the strain distri-
The width of the midheight shear crack was estimated by bution within the CFRP strips and the average strains in the
subtracting half of the element cracking strain (εcr) from the CFRP strips across the critical shear crack at the ultimate
element principal tensile strain (ε1), which was then multi- shear capacity.
plied by the surface elements’ effective length perpendicular One advantage of using CFRP anchors is that they bond well
to the crack (Leffective), as shown in Eq. (1). An in-depth discus- with the CFRP strips, which helps minimize strain concentra-
sion on crack width estimates can be found in Shekarchi20 tions at the interface of the anchor fan and strip. The limited
strain concentrations enabled the CFRP strips to exhibit a
wc = (ε1 – εcr/2) × Leffective (1) uniform strain distribution along the height and across the
width of the 3 in. (76 mm) wide CFRP strips as shown in
The midheight shear crack widths for the uncracked double Fig. 10 for Strip 8 of S-C-VA-HN. However, research14 has
curvature tests are plotted in Fig. 9. The estimated crack shown that wider CFRP strips with one anchor exhibited a less
widths were compared to hand-measured crack widths at uniform strain distribution across the width of the strip. As a
various applied shear loads and locations to validate the result, the average strains in the CFRP strip across the critical
calculations. It was found that the bidirectional layout used crack tend to be lower at the ultimate shear capacity.14,18
on D-U-VF-HA controlled the width of the shear crack For design, a lower-bound average vertical CFRP strain
better at the ultimate shear capacity than the unidirectional (that is, effective CFRP strain) is suggested for estimating
layout used on D-U-VF-HN. The shear crack was restrained the CFRP shear contribution. The average CFRP strain is

ACI Structural Journal/January 2018 287


Table 4—Average and minimum vertical CFRP
strains
Vertical CFRP strain
Test name Average Minimum
S-C-VA-HN 0.0093 0.0072
D-C-VF-HN 0.0079 0.0060
D-U-VF-HN 0.0107 0.0086
D-U-VF-HA 0.0106 0.0096
D-C-VA-HN 0.0076 0.0038
D-U-VA-HA 0.0077 0.0036

determined by averaging the vertical strains in all the CFRP


strips that cross the critical shear crack at the ultimate shear
capacity of the member. The measured average CFRP strip
strains, as well as the minimum CFRP strip strains, at first
strip fracture for all tests are shown in Table 4. The data indi-
cate that the CFRP anchors enabled the CFRP strips to reach
a nearly identical average CFRP strain as the fully wrapped
layout. Specifically, average CFRP strains for D-C-VA-HN
and D-C-VF-HN at first strip fracture reached 0.0076 and
0.0079, respectively.

Retrofitting uncracked and precracked sections


With a preexisting shear crack, the steel reinforcement and
the CFRP strips contributed to the shear capacity at the onset
of loading, as shown in Fig. 11(a) for D-C-VF-HN. In the
uncracked specimen (D-U-VF-HN), the steel contribution Fig. 11—Typical CFRP strain response. (Note: 1 kip =
did not increase until the specimen cracked and the CFRP 4.45 kN.)
did not contribute significantly until the transverse steel
reinforcement yielded (Fig. 11(b)). Specimens D-C-VF-HN Loading conditions
(cracked) and D-U-VF-HN (uncracked), which were nomi- The single and double curvature specimens had a span-
nally identical tests except for precracking, had comparable depth ratio (av/d) of 3.0, although the double curvature speci-
ultimate shear capacities. mens had an effective span-depth ratio (M/Vd) of 1.5 because
However, the precracked member did not reach the same there was a point of inflection at the middle of the test span.
average CFRP strain at failure as its uncracked companion. The effective span-depth ratio implies that the double curva-
In particular, D-C-VF-HN reached an average CFRP strain ture specimens should have exhibited deep beam behavior.
of 0.0079 while D-U-VF-HN reached an average strain of However, the ultimate shear capacities of the single and double
0.0107 at the ultimate shear capacity, which suggests that curvature control specimens were nearly the same. Moreover,
the average CFRP strain across the critical crack may be the strengthened double curvature specimens experienced up
adversely affected by the width of the shear crack prior to the to a 56% increase in shear capacity using CFRP strips and
CFRP installation. The precracked specimens had maximum CFRP anchors despite past studies conducted by Kim et al.10
crack widths of approximately 0.03 in. (0.75 mm) when the and Jirsa et al.14 showing that anchored CFRP strips cannot
CFRP was installed. effectively confine concrete struts or nodal zones. Had the
For design, the differences in the average CFRP strains double curvature specimens exhibited deep beam behavior,
can be accounted for by using a lower-bound effective CFRP the anchored CFRP layouts would not have increased the
strain. Based on Table 4, a reasonable lower-bound effective shear capacity of the members significantly.
strain in the CFRP strips in cracked or uncracked members Based on the similar control capacities, the significant
at the ultimate shear capacity can be taken as 70% of the strength gains with CFRP strengthening, and the fact that
CFRP laminate’s fracture strain (that is, 0.007 for the CFRP the double curvature tests did not form a direct compression
used in this experimental study). The suggested average strut from the applied load to the reaction, it can be concluded
strain is considerably larger than the ACI 440.2R-081 that the double curvature specimens exhibited sectional shear
maximum recommended effective CFRP strain of 0.004 for behavior. Consequently, the material shear contributions from
fully wrapped layouts. It is noted that the suggested lower- the single and double curvature specimens can be directly
bound effective CFRP strain was determined for specimens compared to identify the effects the loading conditions had on
that had a maximum coarse limestone aggregate size of 1 in. the shear behavior of the bent cap girders.
(25 mm), which may influence aggregate interlock prior to The shear contributions of the concrete, steel stirrups, and
fracturing the CFRP strips.1,24 CFRP strips were determined from the measured transverse

288 ACI Structural Journal/January 2018


Fig. 12—Normalized shear contributions, or shear stress coefficients, at cracking and ultimate shear capacity.
steel and CFRP reinforcement strains, constitutive relations,
and vertical equilibrium across the critical shear crack. Using
this approach, the ability to estimate the response of the
material shear contributions depends on the strain measure-
ments close to the shear crack. For instance, the shear contri-
bution of the steel reinforcement may be underestimated
prior to the stirrups yielding if the strain gauges used to esti-
mate the internal steel forces were not at or near the critical
shear crack. However, given the elastic-plastic constitutive
relation of the steel, once the strain gauges recorded strains
beyond yield, their force became constant and equal to the
yield strength. Unlike the shear contribution of the steel, the
CFRP strains were more accurately measured at the crit-
ical crack location using the optical measurement system.
However, the vertical CFRP strip strains, and thus the shear
contribution of the vertical CFRP strips, may have been
marginally underestimated before debonding of the CFRP
strips spread from the crack over the selected gauge length
of the target elements. Nevertheless, once the CFRP strips
debonded beyond the tracked elements, the gauge length of
the elements did not affect the measured strains. In fact, the
CFRP strips debonded rapidly and fully after the transverse
steel reinforcement yielded (that is, at strains around 0.002),
which is a relatively low CFRP strain. The shear contribu-
tion of the transverse CFRP reinforcement was calculated
assuming a uniform strain distribution across the width of
the CFRP strips and using an elastic constitutive relation-
Fig. 13—Comparison of single and double curvature control
ship. The vertical component of the concrete shear contri-
tests. (Note: 1 kip = 4.45 kN.)
bution was estimated from vertical equilibrium, which indi-
cates that the concrete carried the remainder of the applied The shear stress coefficients shown in Fig. 12 suggest that
shear force after the steel and CFRP contributions were the double curvature control tests had higher concrete contri-
subtracted from the applied shear. A summary of the normal- butions but lower steel contributions when compared to the
ized material shear contributions at inclined cracking and the single curvature control test (S-U-VN-HN), even though all
measured ultimate shear capacity (VEXP) is shown in Fig. 12. the tests had similar ultimate shear capacities. This apparent
The effects of the loading conditions were investigated by load shift may be attributed to the fact that the single curva-
studying the control specimens (no CFRP strips) to minimize ture control test had a much shallower shear crack angle
the number of unknown variables. The variation in the shear (25 degrees from horizontal) compared to the double curva-
contribution responses, shown in Fig. 13, can be attributed ture control tests (34 and 31 degrees), which caused the
to differences in the flexural cross-sectional detailing and the critical shear crack in the single curvature tests to cross one
loading conditions. additional stirrup.

ACI Structural Journal/January 2018 289


Similar trends were observed for the strengthened single installation resulted in lower average CFRP strains across
and double curvature specimens, as shown in Fig. 12. On the shear crack.
average, the strengthened double curvature specimens had a 6. At failure, the measured lower-bound average (that is,
63% higher concrete contribution compared to the strength- effective) CFRP strain was approximately 0.007 consid-
ened single curvature specimen. ering both cracked and uncracked specimens; a strain that
Dowel action caused by the cross-sectional detailing of is significantly larger than the maximum permitted effective
the flexural reinforcement can artificially inflate the concrete CFRP strain in the ACI 440.2R-081 design guidelines (that
contribution because the effects are lumped into a single is, 0.004). Additionally, the measured effective CFRP strain
contribution. However, the single curvature specimen had is considerably larger than the effective CFRP strain used in
50% more overall flexural reinforcement compared to the ACI 440.2R-08 for bond-critical CFRP layouts, which illus-
double curvature specimen and yet the single curvature trates that the CFRP material is more effectively used when
concrete contribution was lower. Because the transverse the strips are properly anchored. It should be noted that the
reinforcement detailing was consistent between the single suggested effective strain was determined from data on beams
and double curvature specimens and the specimens with with a maximum coarse limestone aggregate size of 1 in.
the least amount of flexural reinforcement experienced the (25 mm) and may be different for other aggregate types
largest concrete contributions, it is likely that the lower or sizes.
effective span-depth ratio of the double curvature specimens 7. The single and double curvature girders had similar ulti-
improved the concrete contribution. mate shear capacities. However, after investigating the mate-
rial shear contributions, it was observed that the internal force
CONCLUSIONS distributions of the single and double curvature specimens
The objective of the bent cap girder experimental program were different. The observed variances in the load distribution
was to investigate the feasibility of strengthening relatively were the result of differences in the shear crack angles and
wide-webbed reinforced concrete members in shear using the effective span-depth ratios. Moreover, the double curva-
CFRP strips and CFRP anchors. In particular, the experi- ture loading was found to result in lower cracking shear loads
mental program was intended to provide insight into the effi- compared to the single curvature loading.
ciency of anchored CFRP layouts relative to fully wrapped
layouts in uncracked and cracked reinforced concrete AUTHOR BIOS
members subjected to typical loading conditions. ACI member William A. Shekarchi received his PhD from the University
of Texas at Austin, Austin, TX. His research interests include shear behavior
The test results showed that relatively wide-webbed rein- of reinforced concrete elements, structural rehabilitation, and the behavior
forced concrete bent cap girders can be strengthened in shear of fiber-reinforced polymers.
effectively using anchored and fully wrapped CFRP layouts.
ACI member Wassim M. Ghannoum is an Associate Professor at the
The following conclusions were based on findings from the University of Texas at San Antonio, San Antonio, TX. He is the Chair of ACI
experimental study. Committee 369, Seismic Repair and Rehabilitation, and a member of ACI
1. The anchored CFRP strips were able to develop the full Subcommittees 318-R, High-Strength Reinforcement (Structural Concrete
Building Code), and 440-F, FRP-Repair-Strengthening (Fiber-Reinforced
tensile strength of the CFRP strips. In fact, anchored CFRP Polymer Reinforcement). His research interests include extending the life
layouts were found to be as efficient as fully wrapped layouts span and increasing the resilience to damage of concrete structures through
even when the anchors were placed in regions of flexural the application of novel structural materials and retrofit techniques.
tension. However, undesirable strain concentrations may ACI Honorary Member James O. Jirsa holds the Janet S. Cockrell Centen-
develop when anchors are overlapped or close to each other. nial Chair Emeritus at the University of Texas at Austin. He is an ACI
2. To engage the CFRP anchors fully, the CFRP strip must Past President and a member of ACI Committee 318, Structural Concrete
Building Code.
fully debond. Hence, the strain compatibility between the
concrete and the CFRP strips cannot be maintained.
ACKNOWLEDGMENTS
3. The post-cracking stiffness of the members prior to The support of the Texas Department of Transportation for Project 0-6783
strengthening increased after strengthening the members is appreciatively acknowledged. The contents of this paper reflect the views
with CFRP layouts. of the authors, who are responsible for the facts and the accuracy of the data
presented herein. The contents do not necessarily reflect the official view
4. Unidirectional layouts were found to provide larger or policies of the Texas Department of Transportation. This paper does not
shear strength gains compared to bidirectional layouts constitute a standard, specification, or regulation. The authors express their
that had the same amount of CFRP surface area on the thanks to the students, faculty, and staff at Ferguson Structural Engineering
Laboratory for their assistance with the project.
web surface. However, the bidirectional layouts exhibited
increased cracking shear loads and reduced shear crack
NOTATION
widths compared to the unidirectional layouts. Specifically, av = length of constant shear span in test region
the unidirectional CFRP layouts were able to increase the d = section’s effective depth taken from extreme compression fiber
control test shear capacity by as much as 56% while the bidi- to centroid of tension steel
fc′ = specified 28-day concrete strength
rectional layouts increased the control test cracking shear Leffective = elements’ effective length perpendicular to the crack
load by as much as 22%. M = moment in shear span being considered
5. Strengthening precracked members in shear using CFRP V = shear force in shear span being considered
wc = width of principal shear crack
layouts was nearly as effective as strengthening uncracked ε1 = element principal tensile strain
members. The presence of shear cracks prior to the CFRP εcr = element cracking strain

290 ACI Structural Journal/January 2018


ρ = ratio of tensile longitudinal reinforcement area to effective area 12. Kim, Y.; Quinn, K.; Ghannoum, W. M.; and Jirsa, J. O., “Strength-
of concrete ening of Reinforced Concrete T-Beams Using Anchored CFRP Mate-
rials,” ACI Structural Journal, V. 111, No. 5, Sept.-Oct. 2014, p. 1027 doi:
10.14359/51686805
REFERENCES 13. Kim, Y.; Ghannoum, W. M.; and Jirsa, J. O., “Shear Behavior of Full-
1. ACI Committee 440, “Guide for the Design and Construction of Exter- Scale Reinforced Concrete T-Beams Strengthened using CFRP Strips and
nally Bonded FRP Systems for Strengthening Concrete Structures (ACI Anchors,” Construction and Building Materials, V. 94, June 2015, pp. 1-9.
440.2R-08),” American Concrete Institute, Farmington Hills, MI, 2008, 80 pp. 14. Jirsa, J. O.; Ghannoum, W. M.; Kim, C.; Sun, W.; Shekarchi,
2. Khalifa, A., and Nanni, A., “Rehabilitation of Rectangular Simply W.; Alotaibi, N.; Pudleiner, D. K.; Zhu, J.; Liu, S.; and Wang, H., “Use
Supported RC Beams with Shear Deficiencies Using CFRP Composites,” of Carbon Fiber Reinforced Polymer (CFRP) with CFRP Anchors for
Construction and Building Materials, V. 16, No. 3, 2002, pp. 135-146. doi: Shear-Strengthening and Design Recommendations/Quality Control Proce-
10.1016/S0950-0618(02)00002-8 dures for CFRP Anchors,” FHWA/TX-16/0-6783-R1D, Center for Trans-
3. Zhang, Z., and Hsu, C. T. T., “Shear Strengthening of Reinforced portation Research, Austin, TX, 2016, 232 pp.
Concrete Beams using Carbon-Fiber-Reinforced Polymer Laminates,” 15. ACI Committee 318, “Building Code Requirements for Structural
Journal of Composites for Construction, ASCE, V. 9, No. 2, 2005, Concrete (ACI 318-14) and Commentary (ACI 318R-14),” American
pp. 158-169. doi: 10.1061/(ASCE)1090-0268(2005)9:2(158) Concrete Institute, Farmington Hills, MI, 2014, 520 pp.
4. Pellegrino, C., and Modena, C., “Fiber-Reinforced Polymer Shear 16. American Association of State Highway and Transportation Offi-
Strengthening of Reinforced Concrete Beams: Experimental Study and cials, “AASHTO LRFD Bridge Design Specifications,” seventh edition,
Analytical Modeling,” ACI Structural Journal, V. 103, No. 5, Sept.-Oct. Washington, DC, 2014, 2150 pp.
2006, p. 720-728. 17. Sun, W., and Ghannoum, W. M., “Modeling of Anchored CFRP
5. Aridome, Y.; Kanakubo, T.; Furuta, T.; and Matsui, M., “Ductility of Strips Bonded to Concrete,” Construction and Building Materials, V. 85,
T-Shape RC Beams Strengthened by CFRP Sheet U-anchor,” Transactions March 2015, pp. 144-156. doi: 10.1016/j.conbuildmat.2015.03.096
of the Japan Concrete Institute, V. 20, 1998, pp. 117-124. 18. Sun, W.; Jirsa, J. O.; and Ghannoum, W. M., “Behavior of Anchored
6. Ortega, C. A.; Belarbi, A.; and Bae, S. W., “End Anchorage of Exter- Carbon Fiber-Reinforced Polymer Strips Used for Strengthening Concrete
nally Bonded FRP Sheets for the Case of Shear Strengthening of Concrete Structures,” ACI Materials Journal, V. 113, No. 2, Mar.-Apr. 2016,
Girders,” Proceedings of the 9th International Symposium on Fiber Rein- pp. 163-172. doi: 10.14359/51688637
forced Polymer Reinforcement for Concrete Structures (FRPRCS-9), 19. Sokoli, D.; Shekarchi, W. A.; Buenrostro, E.; and Ghannoum, W. M.,
Thomas Telford, London, UK, 2009. “Advancing Behavioral Understanding and Damage Evaluation of Concrete
7. Deifalla, A., and Ghobarah, A., “Strengthening RC T-Beams Subjected Members Using High-Resolution Digital Image Correlation Data,” Earth-
to Combined Torsion and Shear Using FRP Fabrics: Experimental Study,” quakes and Structures, V. 7, No. 5, 2014, pp. 609-626. doi: 10.12989/
Journal of Composites for Construction, ASCE, V. 14, No. 3, 2010, eas.2014.7.5.609
pp. 301-311. doi: 10.1061/(ASCE)CC.1943-5614.0000091 20. Shekarchi, W. A., “Shear Behavior of Reinforced Concrete Bridge
8. Orton, S. L.; Jirsa, J. O.; and Bayrak, O., “Design Consid- Pile Cap Girders Strengthened with Carbon Fiber Reinforced Polymer
erations of Carbon Fiber Anchors,” Journal of Composites for (CFRP) Strips and CFRP Anchors,” PhD dissertation, The University of
Construction, ASCE, V. 12, No. 6, 2008, pp. 608-616. doi: 10.1061/ Texas at Austin, Austin, TX, 2016, 317 pp.
(ASCE)1090-0268(2008)12:6(608) 21. Triantafillou, T. C., and Antonopoulos, C. P., “Design of Concrete
9. Huaco, G. D.; Jirsa, J. O.; and Bayrak, O., “Quality Control Test for Flexural Members Strengthened in Shear with FRP,” Journal of Compos-
Carbon Fiber Reinforced Polymer (CFRP) Anchors for Rehabilitation,” ites for Construction, ASCE, V. 4, No. 4, 2000, pp. 198-205. doi: 10.1061/
Tenth International Symposium on Fiber-Reinforced Polymer Reinforce- (ASCE)1090-0268(2000)4:4(198)
ment for Concrete Structures, SP-275, American Concrete Institute, Farm- 22. Chen, J. F., and Teng, T. G., “Shear Capacity of FRP Strengthened
ington Hills, MI, 2011, pp. 1-18. RC Beams: FRP Debonding,” Construction and Building Materials, V. 17,
10. Kim, Y.; Quinn, K. T.; Satrom, C. N.; Garcia, J.; Sun, W.; Ghan- No. 1, 2003, pp. 27-41. doi: 10.1016/S0950-0618(02)00091-0
noum, W. M.; and Jirsa, J. O., “Shear Strengthening of Reinforced and 23. Bae, S. W., and Belarbi, A., “Behavior of Various Anchorage Systems
Prestressed Concrete Beams Using Carbon Fiber Reinforced Polymer Used for Shear Strengthening of Concrete Structures with Externally
(CFRP) Sheets and Anchors,” FHWA/TX-12/0-6306-1, Center for Trans- Bonded FRP Sheets,” Journal of Bridge Engineering, ASCE, V. 18, No. 9,
portation Research, Austin, TX, 2012, 325 pp. 2013, pp. 837-847. doi: 10.1061/(ASCE)BE.1943-5592.0000420
11. Koutas, L., and Triantafillou, T. C., “Use of Anchors in Shear 24. Vecchio, F. J., and Collins, M. P., “The Modified Compression-Field
Strengthening of Reinforced Concrete T-Beams with FRP,” Journal of Theory for Reinforced Concrete Elements Subjected to Shear,” ACI Journal
Composites for Construction, ASCE, V. 17, No. 1, 2013, pp. 101-107. doi: Proceedings, V. 83, No. 2, Mar.-Apr. 1986, pp. 219-231.
10.1061/(ASCE)CC.1943-5614.0000316

ACI Structural Journal/January 2018 291


NOTES:

292 ACI Structural Journal/January 2018


NOTES:

ACI Structural Journal/January 2018 293


NOTES:

294 ACI Structural Journal/January 2018


NOTES:

ACI Structural Journal/January 2018 295


NOTES:

296 ACI Structural Journal/January 2018


ACI
STRUCTURAL
J O U R N A L
J O U R N

The American Concrete Institute (ACI) is a leading authority and


resource worldwide for the development and distribution of
consensus-based standards and technical resources, educational
programs, and certifications for individuals and organizations involved
in concrete design, construction, and materials, who share
a commitment to pursuing the best use of concrete.

Individuals interested in the activities of ACI are encouraged to


explore the ACI website for membership opportunities, committee
activities, and a wide variety of concrete resources. As a volunteer
member-driven organization, ACI invites partnerships and welcomes
all concrete professionals who wish to be part of a respected,
connected, social group that provides an opportunity for professional
growth, networking, and enjoyment.

You might also like