You are on page 1of 14

LRFD SPECIFICATIONS FOR THE DESIGN CURVED STEEL GIRDER BRIDGES

John M. Kulicki, Ph.D., P.E., President/CEO and Chief Engineer Modjeski and Masters, Inc. 4909 Louise Drive Mechanicsburg, PA 17055 ABSTRACT This paper succinctly outlines the evolution of curved girder design specifications from the 1980 AASHTO Guide Specifications for Horizontally Curved Girders through to the LRFD provisions for curved girder design accepted for inclusion in the AASHTO LRFD Bridge Design Specifications at the June 2004 meeting of the Subcommittee on Bridges and Structures. The multi-path approach to develop the background for these specifications is described, and some of the more significant changes to the design process are outlined. HISTORY The history of specification development related to horizontally curved girder bridges spans over 30 years. During the late 1960s and early 1970s, a group of researchers called The Consortium of University Research Teams, or CURT, developed guidance on the analysis of curved girder bridges and characterizations of the strength and stability of curved girders. This work lead to software products and the design provisions that became codified in the 1980 AASHTO Guide Specifications for Horizontally Curved Highway Bridges (1980 Guide) (1). This Specification was initially produced in the working stress design format, but was quickly converted to load factor design (LFD). In 1993, an updated version of the 1980 Guide was released (1993 Guide) (2). Recognizing the need to update the technology in these earlier design specifications, a major research effort was initiated by the FHWA, which involved both experimental and analytic investigations, of curved steel bridges. The experimental work undertaken in the FHWAs Turner-Fairbank Highway Research Laboratory involved a three-girder single-span structure shown in Figure 1. The experimental program has been widely reported and has been ongoing for over a decade (3,4,5). The FHWAs experimental program was augmented by tests of single girders at several universities, and by large scale finite element analysis with nonlinear materials and geometries (6). Comparisons were made between the results of tests on the curved girder test frame and analytic results. A sample comparison of load deflection curves is shown in Figure 2.

One of the pivotal features of the finite element analysis was its ability to accurately represent the capacities due to local flange buckling and lateral torsional buckling, both in terms of the resistance of the cross-section and in terms of the deflected shape. Figure 3 illustrates a local buckle in the compression flange of a test girder in the Turner-Fairbank test bed. Figure 4 shows the analytic prediction of that local buckle. The agreement is excellent. The ability to analytically simulate the behavior of real components in a curved system made it possible to augment a limited number of experimental tests with literally hundreds of analytic investigations.

Figure 1 Horizontally Curved Girder Test (Courtesy, D. White)

Figure 2 Comparison of Experimental and Analytic Load Deflection Behavior Frame (Courtesy, D. White)

Figure 3 Experimental Flange Local Buckle (Courtesy, D. White)

Figure 4 Analytical Buckle Prediction (Courtesy, D. White)

While the experimental and analytic program was underway, an update to the 1993 Guide was prepared under NCHRP Project 12-38 (7,8). The state of the art of curved girder specifications and a review of advances in understanding the resistance of curved sections, which had taken place in the intervening ten years, resulted in a more modern specification, which was still in the LFD format. The provisions for I-girders retained the w equations introduced in the 1980 Guide, but made significant advances in the recognition of the need to directly interrelate the lateral flange bending stress, or

warping stress, with the vertical bending stress. This was done by subtracting part or all of the lateral flange bending stress from the resistance of the cross-section. The need for additional stud connectors in composite sections due to the radial component of shear between the deck and girders was also recognized. Recognizing the need to include curved girder bridges in the AASHTO LRFD Bridge Design Specifications (AASHTO LRFD) (9), the NCHRP initiated Project 12-52 to develop provisions for inclusion in AASHTO LRFD. This work was done under two phases. Phase I was intended to incorporate the basic provisions of the 2003 Guide Specifications with only minor updating. The literature search conducted under NCHRP Project 12-38 (10) was updated (11). Initial calibration studies determined that the loads and load factors currently in the LRFD Specifications could be retained and used with curved systems (12). Specification provisions were prepared for consideration by AASHTO and companion sample designs for both a curved I-girder bridge and a curved box beam bridge were prepared (13,14,15). The original Scope for Project 12-52 envisioned publishing these initial provisions for use until such time as the FHWA curved girder project produced sufficient results to write an updated specification under Phase 2 of the project. In 2001, the NCHRP Panel, directing Project 12-52, became satisfied that the FHWA project was moving quickly enough that within about two years it would be possible to develop provisions based on that work, completely updating the resistance equations. Rather than publishing design provisions and then totally supercede them in approximately two years, the decision was made to not publish the Phase I provisions and instead work toward a schedule that envisioned the adoption of Phase II provisions in 2004, so that it would be available in ample time for the 2007 FHWA deadline for use of AASHTO LRFD on all Federally-funded projects. Based on the work by the FHWA team, it became evident that the resistance formulations being developed were applicable both to straight and curved systems. A task force was assembled to rewrite and streamline the provisions of LRFD Article 6.10 for I-girders to take advantage of emerging research. These new provisions for straight girders used requirements for lateral torsional buckling, local flange buckling, shear resistance, and shear-moment interaction that were suitable both for straight and curved bridges (6,16,17). The improved and streamlined Article 6.10 of the AASHTO LRFD for straight girders was approved by AASHTO in the 2003 meeting of the Subcommittee on Bridges and Structures. This approval paved the way for a relatively straightforward addition of the requirements for curved girders in that the lateral flange bending stress and the magnification factor were already included. In order to produce the draft LRFD curved girder provisions, all that was necessary was to add the wording needed to separate those provisions which applied solely to straight girders from the rest of Section 6 (18). This was largely a matter of confining the design of curved systems to the use of resistance not to exceed yield resistance. No post-yielding resistance was to be utilized for curved systems. Additional information gleaned from the 2003 Guide Specifications was added to the first four sections of the LRFD Specifications, Introduction, General Features, Loads, and Analysis, respectively. Minor changes related to the rotation of

bearings were added to Section 14, Joints and Bearings, and provisions were added to Section 11 of the LRFD Bridge Construction Specifications (19). OVERVIEW OF DESIGN SPECIFICATION CHANGES Space limitations do not permit a complete summary of all of the changes to the design process for steel curved I- and box girder bridges. Only the more significant provisions are described below. Section 2 General Features There are three primary changes to this section. They are: A requirement for the tracking of additional deformations for straight skewed steel bridges and horizontally curved steel bridges, including the elastic vertical lateral and rotational deflections due to applicable load combinations at the service limit states, girder rotations at the bearings, accumulated over the engineers assumed construction sequence, and the development of camber diagrams satisfying the provisions specified in the steel girder design section, Article 6.7.2. When considering the live load deflection of curved steel box and I-girder sections, the deflection of each individual girder is to be determined based on a system response to adequately take into consideration the system torsional response of the structure. With regard to optional criteria for span to depth ratio, the provisions now require that for curved steel girder systems, the span to depth ratio of each of the girders should not exceed 25 when the specified minimum yield strength of the positive moment section is 50 ksi or less, and the minimum specified yield point in negative moment regions is 70 ksi or less. For other steel girder sections, there is a requirement that the span to depth ratio not exceed the following:
Las 50 = 25 D Fyc

(1)

where: Fyc = specified minimum yield strength of the compression flange (ksi) D = depth of steel girder (ft.) Las = an arc girder length defined as the arc span for simple spans, 0.9 times the arc span for continuous end-spans, and 0.8 times the arc span for continuous interior spans (ft.) During construction, the issues to be investigated include consideration of deflection, strength of steel and concrete, and stability during critical stages of the construction process. Additional guidance as to limit states and load factors are discussed below in relation to Section 3, Loads.

Section 3 Loads and Load Factors

Article 3.4.1, Load Factors and Load Combinations, now contains additional information regarding loads transmitted to a steel superstructure due to the prestressing of precast deck elements. The effect of shrinkage and long-term creep around the shear connectors should also be evaluated to ensure that the composite girder is able to resist the prestressing effects, including the effects of creep and shrinkage, over the life of the structure. The contribution of long-term deformations in closure pours between precast deck panels also needs to be considered. There has been a common practice to age precast deck panels to reduce the shrinkage and this helps to control the transfer to the steel structure. However, even at a delayed age of loading, creep deformations continue to take place and need to be evaluated. Article 3.4.2.1, Evaluation at the Strength Limit State, clarifies the application of load combinations and load factors. In essence, the requirements state that: All applicable strength load combinations in Table 3.4.1-1 are to be modified as specified in Article 3.4.2.1. Strength Load Combinations I, III and V are to be evaluated as necessary during the construction process. Load factors are specified for the weight of the structure and appurtenances. For these loads, identified in the LRFD Specifications as DC and DW, the load factors are not to be taken less than 1.25. Unless the owner chooses a different requirement, the load factor for loads during the construction process for equipment and any dynamic effects associated with the equipment are not to be taken less than 1.5 when applied in Strength Load Combination I. Wind during the construction process may also have to be evaluated under Strength Load Combination III, in which case the load factor for the wind should not be taken less than 1.25 and a load factor of 1.0 may be applied to other loads in that load combination.

Article 3.4.2.2, Evaluation of Deflections at the Service Limit State, provides additional guidance for the calculation of deflections during the construction scenario as follows: Unless project-specific requirements state differently, these investigations are to be based on the Service Load Combination I. For the purpose of deflection calculations, the construction loads, even though temporary, would be considered part of the permanent load of the structure. The transient loads which occur during the construction operations would be considered live load.

Since these deflection evaluations are made during construction, there are no criteria in the Specifications for determining the acceptability of the results. Therefore, the deflections to be allowed have to be included in the contract documents. Article 3.3, Centrifugal Force, was originally written for applications primarily to substructure elements. With the addition of provisions for curved superstructures, the radial force tending to cause an overturning effect on the wheel loads needs to be considered. A distinction has been made between the strength and fatigue limit states. The strength limit state considers the possibility that trucks larger than the design truck are on the structure by adjusting the weight of the design truck using a factor of 4/3, but does not require the simultaneous application of the design live lane load. However, for fatigue evaluation, the cumulative damage factors have already been calculated based on a histogram of vehicles, some of which exceed the design vehicle. In this case, it is not necessary to use the additional 4/3 factor. This is accounted for in the Specifications through the application of Equation 3.6.3-1, shown below.
C=f v gR
2

(2)

where: v f g R = = = = highway design speed (ft./sec.) 4/3 for load combinations other than fatigue and 1.0 for fatigue gravitational acceleration: 32.2 (ft./sec.2) radius of curvature of traffic lane (ft.)

The commentary notes that the centrifugal force causes this overturning effect on the wheel loads because it is applied at the assumed center of mass of the structure, typically taken for design specifications at 6 feet above the roadway. This effect tends to increase the apparent weight of the wheels toward the outside of the bridge and causing an equal unloading of the inside wheels. This will add to the general effective curvature to increase load on the outside girders. It is common to super-elevate the roadway deck on curved structures, and this tends to offset the wheel unbalance caused by the centrifugal force. This offset can also be taken into account. Section 4 Analysis Article 4.6.1.2, Structures Curved in Plan, outlines a series of general requirements for the analysis of this type of structure. The provisions require the use of a rational method for the analysis of the structure, which in some acceptable manner, recognizes that the entire superstructure has to be included for the proper evaluation of force effects. Bearings are an important part of the behavior of a curved system and the boundary conditions have to adequately represent the articulation provided by bearings or other integral connections. The commentary contains additional guidance on these general aspects of analysis indicating, among other things, that the small deflection theory is almost always adequate for the analysis of force effect in curved girders. As stated earlier, it is

sometimes necessary that some recognition of large deflection or second order aspects need to be included in regard to the tendency of curved bottom flanges to deflect outward. This is the situation in which the Specifications generally allow for the use of the moment magnification process in lieu of a second order calculation. The question of whether there is some radius beyond which the curvature effects need not be included has plagued curved girder specification developers throughout the history of these provisions. Article 4.6.1.2.4b, I-girders, and Article 4.6.1.2.4c, Closed Box and Tub Girders, seek to clarify this issue. In the case of the I-girders, the Specifications require that the effect of curvature on the stability of structures be considered regardless of how flat the curve is. There are, however, provisions that provide relief for the need to include the curvature when determining the major axis bending moments and shears in curved systems. Where structures meet the following requirements, a straight girder analysis may be used: The girders are concentric The bearing lines are not skewed more than 10 degrees from radial The stiffness of the girders are similar For all spans, the arc span length divided by the radius of curvature is less than .06 radians. Provisions for determining an appropriate span length are included.

Despite the fact that the curvature may be excluded for structures meeting these requirements for the purpose for determining the major axis bending moments, the requirement to address lateral flange bending effects still must be considered, regardless of the flatness of the curve. Where structures have qualified under the provisions above, it will then be necessary to use some sort of an approximate method similar to one of the M/R methods to make an estimate of the lateral bending from which to determine the flange bending stress. The requirements for closed box and tub girders are similar in intent, but differ in some aspects of the criteria for relief from consideration of the curvature for the determination of the major axis bending moments. Some of the classical methods, such as the V-load method (20) or the M/R method have used straight girder distribution factors in order to have a starting point for the consideration of curvature effects. Similarly, structures qualifying for relief from consideration of the effect of curvature on major axis bending effects, as outlined above, could be designed using the distribution factors of the Specifications. In order to accommodate this, Article 4.6.2.2.1, Application, under beam slab bridges, permits the use of the distribution factors under those specific situations.

Section 6 Steel Design Section 6 contains the provisions for determining the resistance of steel construction, as well as detailing requirements. The highlights of the curved girder additions to this section include the following: No post-yield resistance is relied upon Hybrid design and tension field action is permitted (21,22) There is no curvature reduction in shear resistance or stiffener spacing (21,22) The w equations have been eliminated in favor of the moment amplification factor described below (6)

A step forward was taken by including part or all of the lateral flange bending stress, as was done in the 2003 Guide, except that the lateral flange bending stress was now considered part of the load and added to the load side along with the vertical bending stress. The simplification of utilizing part of the lateral flange bending stress, typically 1/3, was thoroughly evaluated and found to be an acceptable approximation over a wide range of parameters (6). The following abridgement of provisions taken from Article 6.10.8 Flexural ResistanceComposite Sections in Negative Flexure and Noncomposite Sections illustrate the application of the 1/3 rule. At the strength limit state, discretely braced flanges in compression and tension must satisfy the following two criteria, respectively: 1 fbu + fl f Fnc (3) 3 1 fbu + fl f Fnt (4) 3 where: f = fbu = f = Fnc = Fnt = resistance factor for flexure flange stress calculated without consideration of flange lateral bending (ksi) flange lateral bending stress (ksi) nominal flexural resistance of the compression flange (ksi) nominal flexural resistance of the tension flange (ksi)

Where a flange is completely braced as in the case of a composite top flange, the later flange bending stress is not considered. Conversely, the provisions for considering compression flanges before they become fully braced require consideration of 100% of the lateral flange bending stress. Since the finite element analysis used to support the development of design provisions included material and geometric nonlinearities, the implication is that the lateral flange bending stress should be determined based on a second order analysis. Most

designers would consider such a requirement as onerous and impractical. In order to make an acceptable design process, provisions were developed to identify where the second order effect was significant enough to include, in which case a moment amplifier is provided. The moment amplifier is similar in principle to the well-known column moment magnification factors that have been in the bridge specification for 30 years was developed. In this way, almost all practical designs would not require the second order analysis. Additionally, through the use of the magnification factor, it was possible to eliminate the w equations from these new provisions. The basic physical phenomena addressed by the w equations is accounted for through the inclusion of the lateral flange bending stress directly as part of the load magnified, as necessary, when accounting for the tendency of the compression flange of the curved girder to continue to move away from the center of the arc. Therefore, the AASHTO LRFD will permit the flange lateral bending stress, f, to be determined directly from first-order elastic analysis for those discretely braced compression flanges which satisfy the following requirement (6):
Lb 1.2Lp Cb Rb fbu / Fyc

(5)

When Equation 5 is not satisfied, the second-order compression-flange lateral bending stresses may be approximated by amplifying first-order values as follows (6):
0.85 fl = 1 fbu Fcr fl 1 fl 1

(6)

where: Cb = moment gradient modifier fbu = largest value of the compressive stress throughout the unbraced length in the flange under consideration, calculated without consideration of flange lateral bending (ksi) Lb = unbraced length (in.) Lp = a limiting unbraced length Rb = web load-shedding factor fbu = largest value of the compressive stress throughout the unbraced length in the flange under consideration, calculated without consideration of flange lateral bending (ksi) f1 = for use in Equation 6, the maximum first-order lateral bending stress in the compression flange under consideration throughout the unbraced length (ksi) Fcr = elastic lateral torsional buckling stress for the flange under consideration If Equation 5 is not satisfied, the compression-flange lateral bending stresses may also be determined from a second-order elastic analysis of the curved system. The erection and cambering of straight skewed bridges and horizontally curved bridges with or without skewed supports is a more complex problem than generally considered.

In some cases, failure to engineer the erection to achieve the intended final position of the girders, or to properly investigate potential outcomes when detailing to achieve an intended final position of the girders, has resulted in construction delays and claims. It is important that engineers and owners recognize the need for an engineered construction plan and the implied level of checking of shop drawings of girders and cross-frames or diaphragms, processing of RFIs, and field inspection. The final geometry of horizontally curved girders has been a source of construction controversies. Article 6.7.2 requires that, for straight skewed I-girder bridges and horizontally curved I-girder bridges with or without skewed supports, the contract documents should clearly state an intended erected position of the girders and the condition under which that position is to be theoretically achieved. Intended erected positions of I-girders in straight skewed and horizontally curved bridges are defined in the provisions as either girder webs theoretically vertical or plumb, or girder webs outof-plumb. Three common conditions under which these intended erected positions can be theoretically achieved are defined as the no-load condition, the steel dead load condition, or the full dead load condition. The issue of fracture criticality of box girder bridges has been clarified in Article 6.11.5. Unless adequate strength and stability of a damaged structure can be verified by refined analysis, in cross-sections comprised of two box sections, only the bottom flanges in the positive moment regions should be designated as fracture-critical. Cross-sections containing more than two box girder sections should not be considered fracture critical. The criteria for a refined analysis used to demonstrate that part of a structure is not fracture-critical has not yet been codified. Therefore, the loading cases to be studied, location of potential cracks, degree to which the dynamic effects associated with a fracture are included in the analysis, the fineness of models and choice of element type should all be agreed upon by the owner and the engineer. The ability of a particular software product to adequately capture the complexity of the problem should also be considered and the choice of software should also be mutually agreed by the owner and the engineer. Relief from the full factored loads associated with the Strength I Load Combination of Table 3.4.1-1 should be considered, as should the number of loaded design lanes versus the number of striped traffic lanes Remaining Tasks in NCHRP 12-52 The two sample designs completed in Phase I will be revised to utilize the newly accepted provision for curved steel bridges. A comparison of the results obtained using the resistance provisions of the 1993 Guide, the 2003 Guide and the new AASHTO LRFD provisions for 22 real bridges are 10 hypothetical bridges will be completed before publication of this paper. These comparisons are based on using the AASHTO LRFD factored loads so that only the variation in specified resistance will be compared.

ACKNOWLEDGEMENTS At various times, Ian M. Friedland, Scott A. Sabol, and David B. Beal, were Program Officers in charge of NCHRP 12-38. Mr. Myint M. Lwin was the Panel Chairman. Mr. Dann Hall was the Principal Investigator of NCHRP 12-38. Mr. Michael A. Grubb was one of the principal authors of the specification developed under NCHRP 12-38 and NCHRP 12-52. Mr. David Beal is the Senior Project Program Officer in charge of NCHRP 12-52. Mr. Edward P. Wasserman was the Panel Chairman. The work of Drs. Wagdy G. Wassef and Danielle D. Kleinhans, Messrs. Christopher W. Smith and Kevin W. Johns, and Ms. Diane M. Long, all of Modjeski and Masters, Inc., on NCHRP 12-52 is gratefully acknowledged. The FHWA experimental program has been under the direction of Dr. William Wright and was initiated in its early stages under the management of Ms. Sheila Duwadi. The numerical evaluations have been conducted by a number of people, but Dr. Donald White deserves special recognition. Dr. Dennis R. Mertz chaired the task force charged with streamlining of the design process for straight girders. Mr. Edward P. Wasserman, P. E., Bridge and Structures Engineer of the Tennessee DOT is acknowledged for his leadership and personal support of all of the projects cited above, as well as for issuing the challenge to complete all the necessary experimental, analytic and provision development work by 2004. NOTATION Cb = D = Fcr = Fnc = Fnt = Fyc = f = fbu = fbu = f = f 1 = g = Las = Lb = Lp = R = moment gradient modifier depth of steel girder (ft.) elastic lateral torsional buckling stress for the flange under consideration nominal flexural resistance of the flange (ksi) nominal flexural resistance of the flange (ksi) specified minimum yield strength of the compression flange (ksi) 4/3 for load combinations other than fatigue and 1.0 for fatigue largest value of the compressive stress throughout the unbraced length in the flange under consideration, calculated without consideration of flange lateral bending (ksi) flange stress calculated without consideration of flange lateral bending (ksi) flange lateral bending stress (ksi) for use in Equation 6, the maximum first-order lateral bending stress in the compression flange under consideration throughout the unbraced length (ksi) gravitational acceleration: 32.2 (ft./sec.2) an arc girder length defined as the arc span for simple spans, 0.9 times the arc span for continuous end-spans, and 0.8 times the arc span for continuous interior spans (ft.) unbraced length (in.) a limiting unbraced length radius of curvature of traffic lane (ft.)

Rb = web load-shedding factor v = highway design speed (ft./sec.) f = resistance factor for flexure REFERENCES 1 AASHTO (1980). Guide Specifications for Horizontally Curved Highway Bridges. American Association of State Highway and Transportation Officials, Inc., Washington, DC. 2 AASHTO (1993). Guide Specifications for Horizontally Curved Highway Bridges. American Association of State Highway and Transportation Officials, Inc., Washington, DC. 3 Duwadi, S. R., D. H. Hall, J. M. Yadlosky, C. H. Yoo, and A. Zureick. 1995. "FHWACSBRP I-Girder Component Testing." Proceedings of ASCE Structures Congress XIII, Boston, MA, April 2-5, 1995, pp. 1695-1698. 4 Grubb, M. A., and D. H. Hall. 2000. "Philosophy and Design of the I-Girder Bending Component Tests." FHWA Curved Steel Bridge Research Project. Report submitted to the FHWA, Turner-Fairbank Highway Research Center, McLean, VA. 5 White, D. W., S. K. Jung, and C. J. Chang. (2002). Design of Curved Composite Test Bridge, Report to PSI Inc. and FHWA, December, 99 pp. 6 White, D. W., A. H. Zureick, N. P. Phoawanich, and S. K. Jung. (2001). Development of Unified Equations for Design of Curved and Straight Steel Bridge I Girders, Final Report to AISI, PSI, Inc. and FHWA, October 547 pp. 7 Yoo, C. H., D. H. Hall, and S. A. Sabol. (1995). "Improved Design Specifications for Horizontally Curved Steel-Girder Highway Bridges." Proceedings of ASCE Structures Congress XIII, Boston, MA, April 2-5, 1995, pp. 1699-1702. 8 Hall, D. H., M. A. Grubb, and C. H. Yoo. (1999). "Improved Design Specifications for Horizontally Curved Steel Girder Highway Bridges" NCHRP Report 424, Transportation Research Board, Washington, D.C. 9 AASHTO (2004). AASHTO LRFD Bridge Design Specifications. 3rd ed. American Association of State Highway and Transportation Officials, Inc., Washington, DC. 10 Yoo, C. H. (1996). Progress Report on FHWA-CSBRP-Task D. FHWA Contract No. DTFH61-92-C-00136, Auburn University Department of Civil Engineering Interim Report submitted to HDR Engineering, Inc., Pittsburgh, Office, Pittsburgh, PA, August 1996. 11 Yoo, C. H., B. H. Choi. (2000). Literature Search Horizontally Curved Steel Girder Bridges, NCHRP 12-52 Quarterly Report, submitted to NCHRP, TRB, Washington, D. C., March 2000. 12 Nowak, A. S., A. Szwed, and T. V. Galambos. (2001). Calibration of LRFD Design Code for Steel Curved Girder Bridges, NCHRP 12-52. University of Michigan. 13 Modjeski and Masters. (2001). Revised Proposed Draft Specifications. NCHRP 1252 (Phase I) Quarterly Report, submitted to NCHRP, TRB, Washington, D. C., December 2001. 14 Modjeski and Masters. (2001). LRFD Design Example Horizontally Curved Steel I-Girder Bridge. NCHRP 12-52 Quarterly Report, submitted to NCHRP, TRB, Washington, D.C., October 2001.

15 Modjeski and Masters. (2002). LRFD Design Example Horizontally Curved Steel Box Girder Bridges. NCHRP 12-52 Quarterly Report, submitted to NCHRP, TRB, Washington, D. C., January 2002. 16 White, D.W., M. Aydemir, and S. K. Jung. (2004). Shear Strength and Moment Shear Interaction in HPS Hybrid I-Girders, Structural Engineering, Mechanics and Materials Report No. 25, School of Civil and Environmental Engineering, Georgia Institute of Technology, Atlanta, GA, April. 17 White, D.W. and M. A. Grubb. (2004). Comprehensive Update to AASHTO LRFD Provisions for Flexural Design of Bridge I-Girders, Building on the Past: Securing the Future, Structures Congress, ASCE, 8 pp. 18 Modjeski and Masters. (2003/2004). LRFD Specifications for Horizontally Curved Steel Girder Bridges Privilege NCHRP 12-52 Panel Review Document, December 2003, updated as Agenda Item #41, AASHTO Subcommittee on Bridges and Structures, June 2004. 19 AASHTO (1988). LRFD Bridge Construction Specifications and Interim Specifications. First Edition. American Association of State Highway and Transportation Officials, Inc., Washington, DC. 20 U.S. Steel Corporation. (1984). V-load Analysis. ADUSS 88-8535-01, available from the National Steel Bridge Alliance (NSBA), Chicago, IL. Pp. 156. 21 Jung, S.K. and D. W. White. (2003). Shear Strength of Horizontally Curved Steel IGirders Finite Element Studies. Report to PSI Inc. and FHWA, July, 136 pp. 22 Zureick, A.H., White, D.W., Phoawanich, N.P., and Park, J. (2002). Shear Strength of Horizontally Curved Steel I-Girders Experimental Tests, Final Report to PSI Inc. and FHWA, 157 pp.

KEY WORDS Bridge Specifications Curved Steel Girders Erected Condition Fracture Criticality LRFD

You might also like