You are on page 1of 15

C R E E P , SHRINKAGE, AND T H E R M A L E F F E C T S

ON M A S S C O N C R E T E STRUCTURE
By Kevin Z. Truman, 1 David J. Petruska, 2 Associate Members, ASCE,
Downloaded from ascelibrary.org by Universidad Politecnica De Valencia on 06/08/15. Copyright ASCE. For personal use only; all rights reserved.

and C. Dean Norman, 3 Member, ASCE

ABSTRACT: A nonlinear incremental structural analysis including thermal loads


was performed on a lock monolith in order to assess the effects of creep, shrinkage,
adiabatic temperature rise, and construction procedures with respect to tensile stresses
and strains. Mass concrete structures are vulnerable to cracking as a result of high
tensile stresses and strains due to thermal loads, material properties, and mechan-
ical loads. The results of a parametric study considering upper and lower bounds
on adiabatic temperature rise, creep, and shrinkage coupled with the incremental
construction model are presented. Stress and strain history plots are shown and
discussed with respect to typical locations within the monolith. The time history
plots are essential in tracking the effects of the construction sequence. The results
presented are used to give guidance for future analysis and designs of mass con-
crete structures.

INTRODUCTION

The state-of-the-art numerical method is used to study the significance of


thermal loads, creep, and shrinkage on mass concrete structures. An incre-
mental analysis formulation, modeling the construction sequence, is em-
ployed to determine what effect these parameters have on the structural re-
sponse during and after construction. Another major reason for performing
such an analysis is to better assess the effects these parameters have on the
tensile stresses and thus better predict when and where cracking may occur
(Truman et al. 1989; Fehl et al. 1988; Bombich et al. 1987).
Cracking in mass concrete structures has been detected as early as the time
of removal of the forms (ACI: "Prediction" 1971; Bazant and Madsen 1983).
Such cracking at early time is commonly attributed to thermal loads. These
thermal loads arise because the reaction of cement with water is exothermic,
causing a liberation of considerable heat during the curing process. The rate
of heat dissipation, as stated by the laws of heat transfer, is proportional to
the inverse of the square of the least dimension; therefore, thermal loads
tend to be insignificant for small concrete members (Holman 1981). For
large concrete members, the volume to surface area ratio is large, causing
heat to "build up" and create large thermally induced strains.
The mass concrete structure to be analyzed is a two-dimensional section
from the auxiliary lock chamber being constructed at Melvin Price Locks
and Dam on the Mississippi River at Alton, Illinois. The lock chamber, with
construction lifts indicated, is shown in Fig. 1. A time history analysis is
performed that closely models the incremental construction. The finite-ele-
ment code used is ABAQUS ("ABAQUS" 1987). Experimental tests made
'Assoc. Prof., Dept. of Civ. Engrg., Washington Univ., St. Louis, MO 63130-
4899.
2
Grad. Student, Dept. of Civ. Engrg., Washington Univ., St. Louis, MO.
3
Res. Struct. Engrg., Waterways Experiment Station, Vicksburg, MS 39180-0631.
Note. Discussion open until November 1, 1991. To extend the closing date one
month, a written request must be filed with the ASCE Manager of Journals. The
manuscript for this paper was submitted for review and possible publication on May
8, 1990. This paper is part of the Journal of Engineering Mechanics, Vol. 117,
No. 6, June, 1991. ©ASCE, ISSN 0733-9399/91/0006-1274/$1.00 + $.15 per page.
Paper No. 25887.

1274

J. Eng. Mech. 1991.117:1274-1288.


STRONG PILES
(13339. LB/IN VERT.
5.0- (B12.S LB/IN HDRIZ. tt.tSYM . )
WEAK PILES
1.5- (7114.3 LB/IN VEHT.
433.3 LB/IN HORIZ.
Downloaded from ascelibrary.org by Universidad Politecnica De Valencia on 06/08/15. Copyright ASCE. For personal use only; all rights reserved.

I I""

D I R E C T I O N OF HEAT
FLOW I N S L A B

10.0'
J 9—ft ft—ft— fT • -ft ft-- ft ft> -
LIFT 1
THERMAL BOUNDARY
CONDITION: NO HEAT
FLOW ACROSS SURFACE
MECHANICAL BOUNDARY
-12.0-^-8.0^ CONDITION: NO
HORIZ. DISPLACEMENT

FIG. 1. 2-D View of Auxiliary Lock Chamber

on samples of the concrete mix to obtain the heat of hydration and creep
were performed at the U.S. Army, Corp of Engineers Waterways Experi-
ment Station, Vicksburg, Mississippi. Shrinkage effects were determined
through testing performed at the University of Michigan (Hanson 1988). The
U.S. Army Corps of Engineers, Waterways Experiment Station, and Ana-
tech, Inc., Vicksburg, Mississippi, used this data in the development of a
FORTRAN subroutine that models the two- and three-dimensional aging,
creep, and shrinkage for the proposed concrete mix ("Development" 1987;
Bombich et al. 1987).

MODELING

Modeling and computer implementation that represent the physical nature


of the structure and its site environment is extremely important if accurate
results are to be obtained from the analysis (Truman et al. 1989). Since this
is a thermal stress analysis, two analyses are actually performed; the first
being the heat transfer analysis, which calculates the nodal temperatures as
a function of time, and the second being the stress analysis, which uses the
nodal temperatures generated from the thermal analysis to calculate the ther-
mal strains, which are superimposed with the strains caused by gravity loads,
creep, and shrinkage. Various load cases were studied varying the degree
of creep and shrinkage within the structure, shown in Table 1.
The important parameters for a thermal analysis of a mass concrete struc-
ture are convection coefficients, ambient temperature of the air, placement
temperature of the concrete, initial soil temperature, placement rate, adi-
abatic temperature rise, the material properties (conductivity, density, and
specific heat), and the thermal boundary conditions at the plane of sym-
1275

J. Eng. Mech. 1991.117:1274-1288.


TABLE 1. Load Cases Examined to Perform Parametric Study of Creep and
Shrinkage

Load case Creep Shrinkage


Downloaded from ascelibrary.org by Universidad Politecnica De Valencia on 06/08/15. Copyright ASCE. For personal use only; all rights reserved.

d) (2) (3)
l Lower bound" Lower boundb
2 Lower bound Upper boundd
3 Lower bound None
4 Upper boundc Lower bound
5 Upper bound Upper bound
6 None Lower bound
"15% of (5a).
"90% of (4a).
c
110% of (4a).
"61.8% of (5a).

metry. The convection coefficients used in this analysis are based on a forced
convection mode, because as the fluid (air) flows over the surface with some
velocity, heat is carried away, producing a temperature gradient within the
concrete. Therefore, the coefficient is a function of the mean wind velocity.
For external surfaces, the mean wind velocity used is 10 mph. Inside the
culvert and gallery (see Fig. 1), a velocity of 1 mph was used because the
air flow is permissible through the voids at the ends of the monoliths. To
simulate the effect of the forms, a coefficient based on the insulating effect
provided by 3/4 in. plywood is used. This modeling resulted in the follow-
ing convection coefficients. The values of the coefficient are in units of Btu/
day-sq in.-°F. For the exterior walls, the coefficient used was 0.141 with
forms and 0.679 without forms; and for inside the voids, a value of 0.125
with forms was used and 0.422 without forms (Bombich et al. 1987).
The forms are removed two days after the lift is placed, with lift placement
proceeding at five-day intervals. These rates are much faster than what ac-
tually occur on the job site; however, these rates also produce higher thermal
gradients. This is a result of the forms being removed near the time when
the peak temperature within the concrete is obtained. Placing new lifts every
five days results in higher temperatures since the new lift adds more heat to
the previous lift prior to a significant amount of cooling.
The temperature of the convective medium is the mean daily ambient tem-
perature as a function of time, which is representative of the project site
conditions as shown in Fig. 2. All computer analyses were based on a con-
struction starting date of 1 July. In addition, concrete placement temperature
is constant for all lifts, with the temperature being 65° F.
The size of the time step will obviously determine how fast the solution
process proceeds. Larger elements and time step will give a rapid, and thus
less costly, solution; smaller elements and time step will give a more ac-
curate solution. The relationship between the time step and element size
recommended by ABAQUS is

U)
* >(«)**
where At = time step; p = density; C = specific heat; k = thermal con-
1276

J. Eng. Mech. 1991.117:1274-1288.


100 i-

90 -

80 ^
Downloaded from ascelibrary.org by Universidad Politecnica De Valencia on 06/08/15. Copyright ASCE. For personal use only; all rights reserved.

70
F "
— 50 -
UJ
cr
=. SO -
<
cr
UJ 40 -

20 -

10 -

0I 1 1 1 1 1 1 1
0 50 100 150 200 250 300 350 400
ELAPSED TIME FROM 1 JULY (DAYS)

FIG. 2. Mean Dally Ambient Temperature for Alton, Illinois

ductivity; and AL = element size. These restrictions are imposed to ensure


numerical stability and convergence of the solution. Table 2 shows typical
values for the concrete material properties, which were then used to predict
the maximum element size. For a one-quarter-day time step, the element
size must be less than 13.9 in., as given by (1).
The time increments selected are 6 hr for the first two days after a lift is
placed, followed by 12 hr increments for the next three days after a lift is
placed. Analyses were made with the element dimension exceeding the re-
strictions of (1) by approximately 25% in the direction of negligible heat
flow (see Fig. 1). Comparing the results with previous analyses, which did
not violate (1), it was found that this had no significant effect on the results
(Truman et al. 1989). Consequently, in the slab, the element size was in-
creased in the horizontal direction since the heat flow is primarily in the
vertical direction. This allowed a 10% reduction in the number of elements
used in the model.
The actual structure is founded on piles, but because of dewatering of the
coffer dam and the placement of the first lift directly on the soil, soil in-
teraction does occur. To model the insulating effect of the soil, elements
with soil properties are used in the finite-element model, with the depth
being 10 ft. In a-priori analysis showed that an insignificant amount of heat
flowed past this depth (Truman et al. 1991). Obviously, the initial soil tem-
perature is depth dependent, therefore, Fig. 3 shows the temperature distri-

TABLE 2. Thermal Analysis Material Properties


Material property Units Concrete Soil
(1) (2) (3) (4)
Conductivity Btu-in./day-sq in.-°F 2.3494 1.584
Density lb/cu in. 0.0871 0.0758
Specific heat Btu/lb-°F 0.21 0.45

1277

J. Eng. Mech. 1991.117:1274-1288.


Downloaded from ascelibrary.org by Universidad Politecnica De Valencia on 06/08/15. Copyright ASCE. For personal use only; all rights reserved.

en
I
CJ
z
_J
M
o
in
LL.
o

Q-
UJ
a

100 -

120

TEMPERATURE (°F)

FIG. 3. Soil Temperature at Start of Construction

bution used at start of construction. This distribution was obtained by per-


forming a thermal analysis over a time period of one year using only the
soil elements exposed to the mean ambient temperature curve shown in Fig.
2 (Truman et al. 1989).
The heat of hydration as a function of time was obtained by determining
the adiabatic temperature rise through testing of concrete specimens at the
U.S. Army Corps of Engineers Waterways Experiment Station. Fig. 4 shows
the results of such a test. The upper-bound adiabatic curve is for a mix
having a heat generation due to hydration of cement of 70 cal/g; the lower
1278

J. Eng. Mech. 1991.117:1274-1288.


70
UPPER BOUND ADIABATIC CURVE (70 cal/gn)

60 LOWER BOUND ADIABATIC CURVE (53 cal/gn)


Downloaded from ascelibrary.org by Universidad Politecnica De Valencia on 06/08/15. Copyright ASCE. For personal use only; all rights reserved.

40

FIG. 4. Upper and Lower Bound of Adiabatic Temperature Rise

bound is 53 cal/g. The heat of hydration is given by


CpAT
Q= (2)
At
where AT = change in temperature. A user-developed FORTRAN subrou-
tine was used to calculate the heat of hydration as a function of time and to
control these values for the different aged lifts (first lift is five days older
than second lift, etc.), which results from the incremental construction.
Due to symmetry of the structure, a thermal boundary condition must be
invoked along the line of symmetry. For a heat transfer analysis, this con-
dition is that no heat flow is permitted in the horizontal direction across the
surface (see Fig. 1).
The values used in the heat transfer analysis for the material properties
are listed in Table 2. All of the aforementioned properties, physical condi-
tions, and parameters were used within the analyses.
The key parameters for the stress analysis are pile location and stiffness,
concrete-soil interaction, mechanical boundary conditions, gravity loads, and
the concrete elastic material properties and constitutive relations. The piles
were modeled using spring elements. The spring stiffness and locations are
shown in Fig. 1. Until the concrete gains adequate strength, the soil supports
the structure, requiring the soil elements to be included in the stress analysis.
In the stress analysis, a mechanical boundary condition must also be ap-
plied at the centerline. This applied boundary condition is no horizontal dis-
placement permitted along the axis of symmetry. Also, the bottom line of
nodes of the soil elements are not allowed to displace in the horizontal or
vertical directions (see Fig. 1).
Gravity loads are applied one day after a lift is placed. Prior to one day,
a distributed load is applied to the top of the previous lift to simulate the
1279

J. Eng. Mech. 1991.117:1274-1288.


effect of the new lift. This is to prevent large displacements prior to time
of set since no support from the vertical formwork is included in the model.
To simulate the formwork used to support the roof of the voids, the vertical
displacements were forced to be linear between the nodes at the top of the
Downloaded from ascelibrary.org by Universidad Politecnica De Valencia on 06/08/15. Copyright ASCE. For personal use only; all rights reserved.

wall. This constraint was applied for seven days after the roof was placed.
In addition, the program redistributed the loads at those nodes. The forms
on the vertical walls in reality provide some support; however, in the anal-
ysis this support was neglected since it would be difficult to model.
Poisson's ratio and the value of the coefficient of thermal expansion are
assumed to be constant with respect to time. The value of the coefficient of
thermal expansion is 4.5 X 1(T6 in./in.-°F. The value of Poisson's ratio is
0.17. Young's modulus for the soil is 3,000 psi, and Poisson's ratio is 0.35.
Young's modulus for the concrete (in psi) as a function of time t, in days,
is given by (Hanson 1988; Bombich et al. 1987)
E(t) = E0 + £,(1 - e-'"<'-'>) + £ 2 (l - e-»2«-i>) (3a)
where
E0 = 1.488 x 106 (3b)
6
Ei = 1.760 X 10 (3c)
6
E2 = 0.0878 x 10 (3d)
«! = 0.060 (3e)
n2 = 0.883 (3/)
Prior to one day, the value is assumed to be linear, with the one-day value
being 1.488 x 106 psi, and at time zero the value is assumed to be zero.
The type of shrinkage used in the model is autogenous shrinkage, the
internal drying caused by the heat of hydration, which can account for up
to 95% of shrinkage in mass concrete (Bombich et al. 1987). Shrinkage
strains are given by the equation (Hanson 1988; Bombich et al. 1987)
e' = C,(l - es") + C 2 (l - eS1') (4a)
where
C, = 102.5 x 10" 6 (4b)
6
C2 = 72.5 x 10" (4c)
st = -0.150 (4d)
s2 = -0.0226 (4e)
The upper bound was obtained by multiplying (4a) by 1.1 and the lower
bound by multiplying (4a) by 0.9, which was representative of the variation
in the test data.
Creep is a complex phenomenon that occurs in concrete upon loading.
Bazant describes creep as the most uncertain mechanical property of concrete
(Bazant and Madsen 1983). When concrete is loaded, the deformation that
results can be broken into two parts: an immediate deformation and a time-
dependent deformation which begins immediately but continues for years,
known as creep.

1280

J. Eng. Mech. 1991.117:1274-1288.


The creep relation used is similar to the age-adjusted effective modulus
method. The creep relation as a function of time, t, age of loading, T, and
temperature, T, is given by (Hanson 1988; Bombich et al. 1987)
2
Downloaded from ascelibrary.org by Universidad Politecnica De Valencia on 06/08/15. Copyright ASCE. For personal use only; all rights reserved.

J(t>t>T) = 2 Afj,T){l - e""i('~T)) + D(j,m - T) (5a)

where
E(3)
Ai(T,T)=A0ie-"RT i = 1, 2 . (.5b)
.E(J).
E(3)~
D(r,T) = D0e -B/RT (5c)
E(j).
r, = 0.18 (Sd)
r2 = 5.30 (5e)
7
Ao, = 3.33 X 10~ (5/)
7
A02 = 3.16 x 10" (5g)
D 0 = 1.761 X 10~9 (5ft)
E(3) = three day Young's modulus; ZS(T) = Young's modulus as a function
of age of loading; B = an experimentally derived constant; and R = gas
constant. The upper-bound value was obtained by multiplying (5a) by 0.618
and the lower-bound value by multiplying (5a) by 0.150, which represented
the variability of the material and test data. Thus, the strain component can
be calculated by
1 5CT(T)
e(0 = + J(f,T,T) + otAr + e* \dt (6)
E(T) 3T

which are used to find the corresponding stresses.

RESULTS

The results at two nodes, indicated by the circles, for the heat transfer
analysis are shown in Figs. 5 and 6. These plots are typical of what was
found on a lift interface and at the exterior surface. Fig. 5 shows an increase
during the first day due primarily to convection, since the air temperature
is approximately 80° F and placement temperature is 65° F. At five days, a
second increase in temperature occurs from conduction of heat liberated by
the new lift. As would be expected, the upper-bound adiabatic curve pro-
duced higher temperatures. After five days, a slight decrease in temperature
occurs for the lower-bound curve, because the new lift is placed at a tem-
perature of 65° F and enough heat is not liberated initially to produce an
increase as in the upper-bound curve. Again, Fig. 6 shows an initial tem-
perature increase due to convection. The decrease after two days is due to
removal of the forms. The remaining results are due to a convection process
with the nodal temperature remaining at nearly the same temperature as the
surrounding air.

1281

J. Eng. Mech. 1991.117:1274-1288.


Downloaded from ascelibrary.org by Universidad Politecnica De Valencia on 06/08/15. Copyright ASCE. For personal use only; all rights reserved.

0 5 10 15 20 25 30 35
RELATIVE TIME (DAYS)

FIG. 5. Temperature versus Time for Upper and Lower Adlabatic Curve

A user-developed computer program was used to search the generated data


for locations where tensile stresses were present. At locations of tensile stresses,
the upper-bound adiabatic curve produced consistently higher stresses. When
thermal loads were neglected, the resulting stresses were generally com-
pressive. Since the upper-bound adiabatic curve produces higher tempera-
tures, it was used in the creep and shrinkage parametric study. The sign
convention used for the stress analysis is positive stresses are tensile and
negative stresses are compressive. Table 1 shows the load cases examined
to perform the parametric study.
The effect on the horizontal and vertical in-plane stress components when
the shrinkage curves are varied and the creep curve is kept constant is shown
in Figs. 7 and 8. The location of the stress is indicated by the plus sign and

100
UPPER BOUND
95 - (70 cal/gm)
LOWER BOUND
90 (53 cal/gm)

85
l/V~'-v'^7^v___
ao

75

70 -
i
1, 1 , . .
0 5 10 15 20 25 30 35 40 45 50
RELATIVE TIME (DAYS)

FIG. 6. Temperature versus Time for Upper and Lower Adiabatic Curve

1282

J. Eng. Mech. 1991.117:1274-1288.


LOAD CASE 1
LOAD CASE 2
LOAD CASE 3
Downloaded from ascelibrary.org by Universidad Politecnica De Valencia on 06/08/15. Copyright ASCE. For personal use only; all rights reserved.

0 5 10 15 20 25 30 35 40
RELATIVE TIME (DAYS)

FIG. 7. Stress versus Time for Varying Degrees of Shrinkage

corresponds to the location of the Gauss point in the element. The stresses
immediately after placement are compressive as a result of thermal expan-
sion, but internal restraint prevents most of this expansion producing a com-
pressive force on the element. Tensile stresses soon occur as a result of
shrinkage and subsequent cooling. The decreases in stress at five-day inter-
vals is a result of increased bending caused by placing the new lifts. The
pattern exemplified in these two plots is typical of what was observed at
other locations where tensile stresses were investigated. From the results, it
can be noticed that shrinkage can reduce the stresses caused by thermal loads.

400
LOAD CASE 1

i
. LOAD CASE 2
i nan TASF 3

300 -

1I
1
1
200 -

-
100 - / * *
^
' —

-100
10 15 20 25 30 35 40
RELATIVE TIME (DAYS)

FIG. 8. Stress versus Time for Varying Degrees of Shrinkage

1283

J. Eng. Mech. 1991.117:1274-1288.


400 ... . LOAD CASE 1
LOAD CASE 4

300
§ LOAD CASE 6
Downloaded from ascelibrary.org by Universidad Politecnica De Valencia on 06/08/15. Copyright ASCE. For personal use only; all rights reserved.

ir
200

100 /s V
— • , _

— ^ "" \._-
0
N

~\

10 15 20 25 30 40
RELATIVE TIME (DAYS)

FIG. 9. Stress versus Time for Varying Degrees of Creep

This is because for the first few days, thermal loads dominate as evidenced
by the rapid increase in tensile stress at early times as seen in Figs. 7 and
8. At approximately day five, however, thermal loading is decreasing while
shrinkage is increasing, causing the element to contract, thus producing a
gradual decrease in stress. Therefore, overestimating the amount of shrink-
age occurring in the system is not necessarily conservative; and neglecting

400 LOAD CASE 1


LOAD CASE 4
LOAD CASE 6

300

200

100

-100
10 15 20 25 30 35 40
RELATIVE TIME (DAYS)

FIG. 10. Stress versus Time for Varying Degrees of Creep

1284

J. Eng. Mech. 1991.117:1274-1288.


Downloaded from ascelibrary.org by Universidad Politecnica De Valencia on 06/08/15. Copyright ASCE. For personal use only; all rights reserved.

RELATIVE TIME (DAYS)

FIG. 11. Strain versus Time for Various Load Cases

shrinkage from the analysis tends to produce conservative results and re-
duced computing time.
The effect of the same stress locations when creep is varied and shrinkage
is constant is shown in Figs. 9 and 10. From the plots it can be concluded
that creep relaxes the stresses with time. Also, much of the creep effect is
obtained between days 2 and 10. Since both creep and shrinkage effects
produce lower stresses within the tensile regions of the structure, it would

RELATIVE TIME (DAYS)

FIG. 12. Strain versus Time for Various Load Cases

1285

J. Eng. Mech. 1991.117:1274-1288.


be conservative to neglect creep and shrinkage and thus simplify the anal-
ysis.
Strains at these two locations were also investigated because cracking cri-
teria for concrete are usually not based purely on stresses or strains alone,
Downloaded from ascelibrary.org by Universidad Politecnica De Valencia on 06/08/15. Copyright ASCE. For personal use only; all rights reserved.

but on their interaction. Figs. 11 and 12 show the strains plotted as a function
of time. In Figs. 11 and 12, an increase in strains occurs in the first two
days due to thermal loads, followed by a decrease in strains due to shrink-
age. In Fig. 11, another increase is noticed at five days due to the next lift
being placed and thus more heat being generated, resulting in an increase
in the strains. After this time, both figures show a gradual decrease in an
exponential fashion. Fig. 12 also shows an upper band formed by load cases
2 and 5 and a lower band formed by load cases 1 and 4. The upper band
represents lower-bound shrinkage; and the lower band is upper-bound shrink-
age. Thus creep shows little effect on the strains, while shrinkage shows a
significant effect. This is because shrinkage is strain-related, and creep is
stress related.
A value commonly used to assume when cracking may occur in concrete
is 100-150 microstrains. The largest strain shown in the figures is approx-
imately 80% of this value. The largest strain found at any location within
the structure was 93.7 microstrains, for the vertical component in the cham-
ber wall in lift 6 two days after the lift was placed for load case 1.

SUMMARY AND CONCLUSIONS

The results of this study provide a starting point for assessing how these
parameters, in particular adiabatic temperature rise, creep, and shrinkage,
will effect the tensile stresses in an incrementally constructed mass concrete
structure. The objective of this work was to link the computer model with
laboratory data and use the results to accurately predict when and where
cracking might occur and then to minimize the extent of cracking that would
actually occur through the use of proper materials and construction proce-
dures. The tensile stresses and strains examined indicated that cracking would
not be present in the monolith for all the load cases. Such a study is shown
here to be feasible using commercially available finite-element codes. The
parametric analysis showed that the higher the heat of hydration of the ce-
ment, the higher the thermal loads and the resulting stresses. Furthermore,
neglecting thermal loads may result in inadequate placement of reinforce-
ment in areas where a static analysis would indicate compressive stresses.
Neglecting creep and shrinkage in an incremental construction analysis pro-
duces conservative tensile stresses since the addition of shrinkage tends to
reduce the thermal loads and creep relaxes the stresses due to thermal and
gravity loads.

ACKNOWLEDGMENTS

The study conducted at Washington University is a part of broader re-


search effort being performed by the U.S. Army Corps of Engineers, St.
Louis District, and the Waterways Experiment Station. The first and second
writers would like to acknowledge the support provided for this study by the
U.S. Army Corps of Engineers contract DACW39-87-K-0065.
1286

J. Eng. Mech. 1991.117:1274-1288.


APPENDIX I. REFERENCES

"ABAQUS—structural and heat transfer finite element code." (1987). User's man-
ual, Hibbitt, Karlsson, and Sorenson, Inc., Providence, R.I.
Downloaded from ascelibrary.org by Universidad Politecnica De Valencia on 06/08/15. Copyright ASCE. For personal use only; all rights reserved.

Baiant, Z. P., and Madsen, H. O. (1983). "Uncertainty analysis of creep and shrink-
age effects in concrete structures." J. Amer. Concrete Inst. Title No. 80-13, 80(2),
116-127.
Bombich, T., Garner, S., Jones, W., and Norman, C. D. (1987). "Thermal stress
analysis of Mississippi River Lock and Dam 26 (R)." Waterways Experiment Sta-
tion of the U.S. Army Corps of Engineers, Vicksburg, Miss.
"Development of two and three dimensional aging and creep model for concrete."
(1987). Report to Waterways Experiment Station of the U.S. Army Corps of En-
gineers, Anatech, Inc., Vicksburg, Miss.
Fehl, B. D., Normal, C. D., and Truman, K. Z. (1988). "Parameters affecting stresses
in mass concrete structures." Corps of Engineers Structural Engineering Confer-
ence, Vol. 1, 389-399, St. Louis, Mo.
Hanson, W. (1988). "Experimental test data reported to the Waterways Experiment
Station of the U.S. Army Corps of Engineers." Report, Vicksburg, Miss.
Holman, J. P. (1981). Heat transfer, 5th Ed., McGraw-Hill Book Co., Inc., New
York, N.Y.
"Prediction of creep, shrinkage, and temperature effects in concrete." (1971). ACI
Publication SP 27-3; Designing for effects of creep, shrinkage, and temperature
in concrete structures, American Concrete Institute, Detroit, Mich., 51-93.
Truman, K. Z., Petruska, D. J., and Ferhi, A. (1989). "Evaluation of thermal and
incremental construction effects for monolith AL-3 and AL-5 of the Auxiliary Lock
and Dam No. 26 (Replacement)." Struct. Engrg. Report No. 84, Washington Univ.,
St. Louis, Mo.
Truman, K. Z., Petruska, D. J., Ferhi, A., and Fehl, B. (1991). "Nonlinear, incre-
mental analysis of a mass concrete lock monolith." J. of Struct. Engrg., ASCE,
New York, N.Y., 117(6), 1834-1851.

APPENDIX II. NOTATION

The following symbols are used in this paper:

•01»A>2 = experimentally obtained constant;


B = experimentally obtained constant;
C = specific heat;
*-\,Ci = experimentally obtained constant;
Do = experimentally obtained constant;
E = Young's modulus;
h = convection coefficient;
J = creep relation;
k = thermal conductivity;
Q = heat of hydration;
R = gas constant;
r\,r2 = experimentally obtained constant;
S\,S2 = experimentally obtained constant;
T = temperature;
t = time;
a = coefficient of thermal expansion;
AL = element dimension;
AT = change in temperature;
At = change in time;
€ = strain;
1287

J. Eng. Mech. 1991.117:1274-1288.


e shrinkage strains;
P density;
a stress; and
age of loading.
Downloaded from ascelibrary.org by Universidad Politecnica De Valencia on 06/08/15. Copyright ASCE. For personal use only; all rights reserved.

APPENDIX III. CONVERSION TO SI UNITS

To Convert To Multiply by

Btu/day-sq in. cal/day-cm2 39.06


op °C [subtract 32 and multiply by 5/9]
ft m 0.305
in. cm 2.54
lb, N 4.45
lbn. kg 0.453
lb/sq ft Pa 47.9
lb/sq in. KPa 6.9
mph m/s 5.7924
sq in. cm2 6.45

1288

J. Eng. Mech. 1991.117:1274-1288.

You might also like