You are on page 1of 118

THE NATIONAL ACADEMIES PRESS

This PDF is available at http://nap.edu/24813 SHARE


   

Strand Debonding for Pretensioned Girders (2017)

DETAILS

102 pages | 8.5 x 11 | PAPERBACK


ISBN 978-0-309-44640-2 | DOI 10.17226/24813

CONTRIBUTORS

GET THIS BOOK Bahram M. Shahrooz, Richard A. Miller, Kent A. Harries, Qiang Yu, and Henry G.
Russell; National Cooperative Highway Research Program; Transportation Research
Board; National Academies of Sciences, Engineering, and Medicine
FIND RELATED TITLES

SUGGESTED CITATION

National Academies of Sciences, Engineering, and Medicine 2017. Strand


Debonding for Pretensioned Girders. Washington, DC: The National Academies
Press. https://doi.org/10.17226/24813.


Visit the National Academies Press at NAP.edu and login or register to get:

– Access to free PDF downloads of thousands of scientific reports


– 10% off the price of print titles
– Email or social media notifications of new titles related to your interests
– Special offers and discounts

Distribution, posting, or copying of this PDF is strictly prohibited without written permission of the National Academies Press.
(Request Permission) Unless otherwise indicated, all materials in this PDF are copyrighted by the National Academy of Sciences.

Copyright © National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

N AT I O N A L C O O P E R AT I V E H I G H W AY R E S E A R C H P R O G R A M

NCHRP RESEARCH REPORT 849


Strand Debonding
for Pretensioned Girders

Bahram M. Shahrooz
Richard A. Miller
University of Cincinnati
Cincinnati, OH

Kent A. Harries
Qiang Yu
University of Pittsburgh
Pittsburgh, PA

Henry G. Russell
Henry G. Russell, Inc.
Glenview, IL

Subscriber Categories
Bridges and Other Structures

Research sponsored by the American Association of State Highway and Transportation Officials
in cooperation with the Federal Highway Administration

2017

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

NATIONAL COOPERATIVE HIGHWAY NCHRP RESEARCH REPORT 849


RESEARCH PROGRAM
Systematic, well-designed research is the most effective way to solve Project 12-91
many problems facing highway administrators and engineers. Often, ISSN 2572-3766 (Print)
highway problems are of local interest and can best be studied by ISSN 2572-3774 (Online)
highway departments individually or in cooperation with their state ISBN 978-0-309-44640-2
universities and others. However, the accelerating growth of highway Library of Congress Control Number 2017941011
transportation results in increasingly complex problems of wide inter-
© 2017 National Academy of Sciences. All rights reserved.
est to highway authorities. These problems are best studied through a
coordinated program of cooperative research.
Recognizing this need, the leadership of the American Association
of State Highway and Transportation Officials (AASHTO) in 1962 ini- COPYRIGHT INFORMATION
tiated an objective national highway research program using modern Authors herein are responsible for the authenticity of their materials and for obtaining
scientific techniques—the National Cooperative Highway Research written permissions from publishers or persons who own the copyright to any previously
Program (NCHRP). NCHRP is supported on a continuing basis by published or copyrighted material used herein.
funds from participating member states of AASHTO and receives the Cooperative Research Programs (CRP) grants permission to reproduce material in this
full cooperation and support of the Federal Highway Administration, publication for classroom and not-for-profit purposes. Permission is given with the
United States Department of Transportation. understanding that none of the material will be used to imply TRB, AASHTO, FAA, FHWA,
FMCSA, FRA, FTA, Office of the Assistant Secretary for Research and Technology, PHMSA,
The Transportation Research Board (TRB) of the National Academies
or TDC endorsement of a particular product, method, or practice. It is expected that those
of Sciences, Engineering, and Medicine was requested by AASHTO to reproducing the material in this document for educational and not-for-profit uses will give
administer the research program because of TRB’s recognized objectivity appropriate acknowledgment of the source of any reprinted or reproduced material. For
and understanding of modern research practices. TRB is uniquely suited other uses of the material, request permission from CRP.
for this purpose for many reasons: TRB maintains an extensive com-
mittee structure from which authorities on any highway transportation
subject may be drawn; TRB possesses avenues of communications and NOTICE
cooperation with federal, state, and local governmental agencies, univer-
The research report was reviewed by the technical panel and accepted for publication
sities, and industry; TRB’s relationship to the Academies is an insurance
according to procedures established and overseen by the Transportation Research Board
of objectivity; and TRB maintains a full-time staff of specialists in high- and approved by the National Academies of Sciences, Engineering, and Medicine.
way transportation matters to bring the findings of research directly to
The opinions and conclusions expressed or implied in this report are those of the
those in a position to use them. researchers who performed the research and are not necessarily those of the Transportation
The program is developed on the basis of research needs identified Research Board; the National Academies of Sciences, Engineering, and Medicine; or the
by chief administrators and other staff of the highway and transporta- program sponsors.
tion departments and by committees of AASHTO. Topics of the highest The Transportation Research Board; the National Academies of Sciences, Engineering,
merit are selected by the AASHTO Standing Committee on Research and Medicine; and the sponsors of the National Cooperative Highway Research Program
(SCOR), and each year SCOR’s recommendations are proposed to the do not endorse products or manufacturers. Trade or manufacturers’ names appear herein
AASHTO Board of Directors and the Academies. Research projects solely because they are considered essential to the object of the report.
to address these topics are defined by NCHRP, and qualified research
agencies are selected from submitted proposals. Administration and
surveillance of research contracts are the responsibilities of the Acad-
emies and TRB.
The needs for highway research are many, and NCHRP can make
significant contributions to solving highway transportation problems
of mutual concern to many responsible groups. The program, however,
is intended to complement, rather than to substitute for or duplicate,
other highway research programs.

Published research reports of the

NATIONAL COOPERATIVE HIGHWAY RESEARCH PROGRAM


are available from

Transportation Research Board


Business Office
500 Fifth Street, NW
Washington, DC 20001

and can be ordered through the Internet by going to


http://www.national-academies.org
and then searching for TRB
Printed in the United States of America

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

The National Academy of Sciences was established in 1863 by an Act of Congress, signed by President Lincoln, as a private, non-
governmental institution to advise the nation on issues related to science and technology. Members are elected by their peers for
outstanding contributions to research. Dr. Marcia McNutt is president.

The National Academy of Engineering was established in 1964 under the charter of the National Academy of Sciences to bring the
practices of engineering to advising the nation. Members are elected by their peers for extraordinary contributions to engineering.
Dr. C. D. Mote, Jr., is president.

The National Academy of Medicine (formerly the Institute of Medicine) was established in 1970 under the charter of the National
Academy of Sciences to advise the nation on medical and health issues. Members are elected by their peers for distinguished contributions
to medicine and health. Dr. Victor J. Dzau is president.

The three Academies work together as the National Academies of Sciences, Engineering, and Medicine to provide independent,
objective analysis and advice to the nation and conduct other activities to solve complex problems and inform public policy decisions.
The Academies also encourage education and research, recognize outstanding contributions to knowledge, and increase public
understanding in matters of science, engineering, and medicine.

Learn more about the National Academies of Sciences, Engineering, and Medicine at www.national-academies.org.

The Transportation Research Board is one of seven major programs of the National Academies of Sciences, Engineering, and Medicine.
The mission of the Transportation Research Board is to increase the benefits that transportation contributes to society by providing
leadership in transportation innovation and progress through research and information exchange, conducted within a setting that is
objective, interdisciplinary, and multimodal. The Board’s varied committees, task forces, and panels annually engage about 7,000
engineers, scientists, and other transportation researchers and practitioners from the public and private sectors and academia, all of
whom contribute their expertise in the public interest. The program is supported by state transportation departments, federal agencies
including the component administrations of the U.S. Department of Transportation, and other organizations and individuals interested
in the development of transportation.

Learn more about the Transportation Research Board at www.TRB.org.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

COOPERATIVE RESEARCH PROGRAMS

CRP STAFF FOR NCHRP RESEARCH REPORT 849


Christopher J. Hedges, Director, Cooperative Research Programs
Lori L. Sundstrom, Deputy Director, Cooperative Research Programs
Waseem Dekelbab, Senior Program Officer
Gary A. Jenkins, Senior Program Assistant
Eileen P. Delaney, Director of Publications
Scott E. Hitchcock, Editor

NCHRP PROJECT 12-91 PANEL


Field of Design—Area of Bridges
Bijan Khaleghi, Washington State DOT, Tumwater, WA (Chair)
Upul Bandara Attanayake, Western Michigan University, Kalamazoo, MI
Fouad A. H. Jaber, Nebraska Department of Roads, Lincoln, NE
Edmund H. Newton, Retired, South Dartmouth, MA
William N. Nickas, Precast/Prestressed Concrete Institute, Tallahassee, FL
Taya Retterer, Texas DOT, Austin, TX
Joshua James Sletten, Parsons Brinckerhoff, Murray, UT
Julius F. J. Volgyi, Jr., Retired, Richmond, VA
Benjamin A. Graybeal, FHWA Liaison
Stephen F. Maher, TRB Liaison

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

FOREWORD

By Waseem Dekelbab
Staff Officer
Transportation Research Board

This report provides proposed revisions to the current debonding provisions found within
the AASHTO LRFD Bridge Design Specifications with detailed examples of the application
of the proposed revisions. The proposed revisions are based on comprehensive analytical
and testing programs for investigating the effects of end anchorages, beam sections, end-
diaphragm details, concrete strengths up to 15 ksi, and strand sizes. The material in this
report will be of immediate interest to highway design engineers.

Strand debonding is an alternative for reducing stresses in the end regions of pretensioned
concrete beams. The AASHTO LRFD Bridge Design Specifications currently limit the amount
of partial debonding to twenty-five percent of the total strand area within a pretensioned
girder. The limit was imposed in recognition of the detrimental effects that excessive debond-
ing can have on shear performance. Nevertheless, several states allow significantly higher
percentages of debonding (up to seventy-five percent) to be used routinely in design. These
higher percentages are based on successful past practices that have not been challenged. It is
clear that unless the experimental evidence provides sufficient clarity, a consensus agreement
among bridge owners will continue to be difficult to achieve.
A comprehensive study of partial debonding effects on the performance of pretensioned
girders was therefore needed. Review of the various debonding practices used throughout
the United States and existing test data allowed focused experimental research on the critical
parameters. The research was designed to produce definitive recommendations regarding
the number and configuration of debonded strands within commonly used cross-sectional
shapes (i.e., I-, U-, and box beams). Final statements regarding the integral role of strand
anchorage in the service and strength performance of pretensioned beams should be well-
substantiated and ultimately highlight the importance of a unified approach to strand
debonding.
Research was performed under NCHRP Project 12-91 by the University of Cincinnati to
develop a proposed revision to the current debonding provisions found within the AASHTO
LRFD Bridge Design Specifications and the AASHTO LRFD Bridge Construction Specifica-
tions. The proposed revisions consider service and strength limit states for strand debonding
within pretensioned flexural superstructure members (i.e., I-, U-, and box beams).
A number of deliverables, provided as appendices, are not published but are available on
the TRB project website (www.trb.org, search for “NCHRP 12-91”). These appendices are
titled as follows:

• Appendix A—Survey
• Appendix B—Design Case Studies

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

• Appendix C—Finite Element Modeling


• Appendix D—Summaries of Individual FEM Simulations
• Appendix E—Specimen Details and Fabrication Photographs
• Appendix F—Material Properties and Mix Designs
• Appendix G—Internal and External Instrumentation
• Appendix H—Overview of Design Calculations
• Appendix I—AASHTO Bridge Committee Agenda Item
• Appendix J—Example of Proposed Changes

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

CONTENTS

xi Notations
1 Chapter 1  Background
1 1.1 Introduction
2 1.2  Objectives of Research Program
2 1.3  Review of State of the Art and Practice
2 1.3.1  Longitudinal Reinforcement Requirements Associated
with Shear Capacity
3 1.3.1.1  AASHTO LRFD Article 5.8.3.5:
Longitudinal Reinforcement
4 1.3.2  Extant Experimental Studies
7 1.3.2.1  Synthesis of Past Experimental Studies
7 1.3.3  Past Analytical Studies
8 1.3.3.1  Synthesis of Past Analytical Studies
8 1.4  Current Practice for Debonding Strand
8 1.4.1  State Amended Specifications
9 1.4.2  Survey of Current Practice
11 Chapter 2  Analytical Research Approach and Findings
11 2.1  Research Approach
11 2.2  Evaluation of Current AASHTO Limits on Strand Debonding
12 2.2.1  Shahawy et al. (1993)
15 2.2.2  Russell et al. (2003)
15 2.3  Design Case Studies
15 2.3.1  Determination of Maximum Girder Span
16 2.3.1.1  STRENGTH I
16 2.3.1.2  SERVICE I (AASHTO LRFD Articles 3.4.1 and 5.9.4.2.1)
17 2.3.1.3  SERVICE III (AASHTO LRFD Articles 3.4.1 and 5.9.4.2.2)
17 2.3.1.4  Minimum Girder Span Length
17 2.3.2  Debonding Ratio
18 2.3.3  Summary of Design Parameter Study
21 2.4  Finite Element Method Modeling
22 2.4.1  Development and Validation of 3D-FEM Model
22 2.4.1.1  Material Models
23 2.4.1.2  Structural Modeling
23 2.4.2  FEM Parametric Study
24 2.4.2.1  Modeling Parameters
24 2.4.2.1.1  Concrete Strength
24 2.4.2.1.2  Prestressing Strand
25 2.4.2.1.3  Partial Debonding
25 2.4.2.1.4  Shear Reinforcement
25 2.4.2.1.5  Boundary Conditions
25 2.4.2.1.6  Applied Loads

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

26 2.4.2.1.7  Modeling Steps


26 2.4.2.1.8  Model Conventions
26 2.4.2.2  Simulation Summary
26 2.4.2.2.1  Prestress Transfer
29 2.4.2.2.2  SERVICE I Limit State
29 2.4.2.2.3  STRENGTH I Limit State
30 2.4.2.2.4  Ultimate Capacity
31 2.4.2.2.5  Performance of Different Girder Cross Sections
31 2.5  Strut-and-Tie Modeling of End Regions
32 2.5.1  Motivating Example
32 2.5.2  STM of Prestressed Girder End Region
34 2.5.3  Illustrative Examples
42 2.5.4  Validation by Experimental Results
43 2.5.5  Summary of End Region Behavior
43 2.5.6  Extension of Bulbed Girder Results and Discussion
of Other Girder Shapes
44 2.6  Evaluation of the Effects of Skewed Girder Ends
44 2.6.1  Mechanistic Modeling
46 2.6.2  Washington State Girder Examples
46 2.6.2.1  Nonlinear Finite Element Modeling of Girders G2 and G5
50 2.7  Introduction of Debonded Strands at One Section
52 Chapter 3  Experimental Research Approach, Findings,
and Associated Analytical Simulations
52 3.1  Research Approach
52 3.2  Design and Detailing of Test Specimens
53 3.3  Material Properties
54 3.4  Transfer Length
59 3.5  Testing Program
59 3.5.1  Test Setup
61 3.5.2 Instrumentation
62 3.5.3  Test Results and Discussions
62 3.5.3.1  Capacity, Stiffness, and Failure Mode
64 3.5.3.2  Crack Patterns
70 3.5.3.3  Shear Deformation
70 3.5.3.4  Shear Resistance from Transverse Reinforcement
72 3.5.3.5  Apparent Strand Slip
76 3.5.3.6  Contribution of Longitudinal Reinforcement
76 3.6 Summary
78 3.7  Modeling of Test Specimens
78 3.7.1  FEM Simulation of Test Girders
78 3.7.2  Utilization of Calibrated Analytical FEM Platform
78 3.7.2.1  Transfer Length
78 3.7.2.2  Further Evaluation of Longitudinal Strains
at Release in Texas U-40
86 3.7.2.3  STM Simulation of Test Girders
89 3.8  Web Cracking
90 3.8.1  Calculation of Principal Tensile Stress
90 3.8.2  AASHTO LRFD Specifications Article 5.8.5
92 3.8.3  AASHTO LRFD Specifications Article 5.8.3.4.3
94 3.8.4  Evaluation of Data for AASHTO BI-36 Test Girder

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

96 Chapter 4  Conclusions and Suggestions


96 4.1 Conclusions
96 4.2  Suggested Detailing Guidelines for Prestressed Girders
Having Debonded Strands
98 4.3  Suggestions Regarding Transverse Tension Ties
at STRENGTH I Ultimate Limit State
99 4.4  Web Shear Cracking
100 4.5  Suggestions for Future Research
101 References
103 Appendices

Note: Photographs, figures, and tables in this report may have been converted from color to grayscale for printing.
The electronic version of the report (posted on the web at www.trb.org) retains the color versions.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

N ot a t i ons

a = shear span (in.)


A = Ramberg-Osgood function parameter
Abulb = area of the bottom bulb (flange)
Abulb,transformed = area of the transformed bulb section
Ac = gross area of precast girder alone (in.2)
Ae,g = “effective” gross area of the concrete section (in.2)
Ae,transformed = “effective” area of the transformed section (in.2)
Ag = gross area of precast girder alone (in.2)
Agirder = area of precast girder alone (in.2)
Anc = area of the non-composite section
Ap = area of one prestressing steel strand (in.2)
Aps = area of prestressing steel (in.2); area of one prestressing strand (in.2)
As = area of nonprestressed tension reinforcement (in.2)
Agross = gross cross-sectional area
Atransformed = area of the transformed section
Av = area of shear reinforcement (in.2)
B = Ramberg-Osgood function parameter
B2 = girder dimension (PCI 2011)
B3 = girder dimension (PCI 2011)
bb = bearing width (in.)
bv = width of web (in.)
C = Ramberg-Osgood function parameter
cb = (bb /2)(1 – nf /Nw)(required location of resultant of portion of shear force dis-
tributed to flange outstands to ensure uniform bearing pressure across the
bearing pad)
D5 = girder dimension (PCI 2011)
D6 = girder dimension (PCI 2011)
DC = weight of supported structure (kip)
DW = superimposed dead load (kip)
d = distance from compressive face to centroid of reinforcement (in.)
db = nominal diameter of reinforcing bar, wire, or prestressing strand (in.)
dc = distance to critical section (in.)
de = effective depth from extreme compression fiber to the centroid of the tensile
force in the tensile reinforcement (in.)
dr = debonding ratio
dv = effective shear depth (in.)
Ec = modulus of elasticity for concrete (ksi)
Ep = modulus of elasticity for prestressing strands (ksi)
Es = modulus of elasticity for mild steel (ksi)
Esh = modulus of elasticity for strain hardening portion of stress-strain curve (ksi)

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

e = eccentricity of the prestressing strands


eg = eccentricity of the prestressing strands measured to the centroid of the gross
concrete section (in.)
egirder = average eccentricity of precast girder alone (in.)
etransformed = eccentricity of the prestressing strands measured to the centroid of the trans-
formed section (in.)
fc = compressive stress in concrete (ksi)
fc,transformed = concrete stress calculated using transformed section properties
fc,g = concrete stress calculated using gross section properties
f c′ = specified compressive strength of concrete for use in design (ksi)
f c′,girder = specified compressive strength of girder concrete for use in design (ksi)
f c′,slab = specified compressive strength of slab concrete for use in design (ksi)
f c′i = concrete compressive strength at prestress transfer (ksi)
fpc = normal stress in the web
fpi = initial prestress at transfer (ksi)
fpLT = average stress in prestressing steel after long-term losses (ksi)
fpe = effective stress in the prestressing steel after losses (ksi)
fps = average stress in prestressing steel at nominal flexural resistance (ksi)
fpu = ultimate strength of prestressing strand (ksi)
fss = steel stress (ksi)
ft = tensile stress in concrete (ksi)
fu = ultimate strength of reinforcing bar (ksi)
fvy = yield strength of transverse reinforcement
fv = stress in shear reinforcement
fy = specified minimum yield strength of reinforcing bar (ksi); yield strength of
reinforcing bar (ksi)
h = overall depth of a member (in.)
H = overall depth of a member (in.)
hb = depth of bulb or bottom flange (in.)
Ic = moment of inertia of the composite section (in.4)
Ig = moment of inertia of the gross concrete section (in.4)
IM = impact loads (kip)
Inc = moment of inertia of the non-composite section (in.4)
Itransformed = moment of inertia of the transformed section (in.4)
K1 = correction factor for source of aggregate
ld = development length (in.)
ldebond = debonded length
L = span length (in. or ft)
Lbearing = length of bearing pad
Li = development length of bonded strand at i for beam with skew angle (in.)
LL = live load (kip)
LLlane = live load due to distributed lane load (ksf )
LLtruck = live load due to HS-93 axel loads (kip)
Lmin = minimum girder span based on the least of LSERVICE I, LSERVICE III, and LSTRENGTH I (ft)
LSERVICE I = maximum girder span as determined by Service I load combination (ft)
LSERVICE III = maximum girder span as determined by Service III load combination (ft)
LSTRENGTH I = maximum girder span as determined by Strength I load combination (ft)
Lt = strand transfer length (ft or in.)
M = bending moment
Mbarrier wall = total bending moment due to barrier wall (kip-in.)
MDC = total bending moment due to weight of supported structure
Mdnc = moment due to external loads applied only to the non-composite section (kip-in.)
MDW = total bending moment due to superimposed dead load (kip-in.)
Mgirder = total bending moment due to weight of girder (kip-in.)

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

ML = moment due to external loads applied only to the composite section (kip-in.)
MLL = total bending moment due to live load (kip-in.)
Mn = nominal flexural resistance (kip-in.)
Mo = moment about point o
Mslab = total bending moment due to weight of slab (kip-in.)
Msw = moment due to the self-weight of the concrete section only (kip-in.)
Mtotal = total bending moment applied in an experimental setting (kip-in.)
Mu = applied factored bending moment at section (kip-in.)
MWS = total bending moment due to wind (kip-in.)
N = number of strands
Nmax = maximum number of strands that may be located in the bulb or lower flange
Nt = property of the section geometry and concrete strength at prestress transfer only
Nu = applied factored axial force at section taken as positive if tensile (kip)
Nw = number of bonded strands at a section
nf = number of bonded strands in one side of the outer portion of bulb. The outer
portion of bulb is defined as that extending beyond projection of web width, B3.
Strands aligned with the edge of web are assumed to fall in the outer portion
of bulb.
ni = number of bonded strands in vertical group i
P = concentrated applied load (kip); effective prestressing force (kip)
Ppe = effective prestressing force after all losses (kip)
Pi = force associated with each vertical group of ni bonded strands at location i (kip)
Q = force effect in associated units
Qc = first moment of inertia of the area above y taken about the neutral axis of the
noncomposite section (in.3)
Qnc = first moment of inertia of the area above y taken about the neutral axis of the
composite section (in.3)
S = girder spacing (ft)
Sbottom of girder = section modulus to bottom of girder alone (in.3)
Sbottom of girder, composite = section modulus to bottom of composite girder (in.3)
Stop of girder = section modulus to top of girder alone (in.3)
Stop of girder, composite = section modulus to top of composite girder (in.3)
s = average spacing between mild shear reinforcement
T = tensile force in the longitudinal reinforcement (kip)
t = composite slab thickness (in.); tension factor for horizontal tie caused by shear
strut at support
tf = flange thickness
ts = thickness of slab
tw = thickness of web
v = shear stress in concrete
Vc = nominal shear resistance provided by tensile stresses in the concrete (kip)
Vci = nominal shear resistance provided by concrete when inclined cracking results
from combined shear and moment (kip)
Vcw = nominal shear resistance provided by concrete when inclined cracking results
from excessive principal tensions in web (kip)
Vdesign = design shear
Vdnc = shear force applied to the non-composite section only (kip)
Vexp = experimentally reported shear values at failure (kip)
VL = shear force applied to the composite section (kip)
Vn = nominal shear resistance of the section considered (kip)
Vp = component in the direction of the applied shear of the effective prestressing
force (kip)
Vs = shear resistance provided by shear reinforcement (kip)
Vshear = total shear resistance (kip)

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

VT@dc = tensile force in longitudinal steel due to shear at the critical section (kip)
VT@support = tensile force in longitudinal steel due to shear at the support (kip)
Vtotal = total shear load applied in an experimental setting (kip)
Vu = applied factored shear force at section (kip)
wc = unit weight of concrete (kcf)
wmax = widest crack observed during testing of specimen
xf = distance from end of beam where the entire girder cross section resists the pre-
stressing force
xp = horizontal distance to girder centerline of centroid of nf strands in outer portion
of bulb
y = vertical distance from the bottom of the section to the point where the stress is
calculated
ybc = vertical distance from the bottom of the section to the composite neutral axis
ybnc = vertical distance from the bottom of the section to the non-composite neutral axis
yc,g = distance to the centroid of the gross concrete section from the bottom of the girder
yc,transformed = distance to the centroid of the transformed section from the bottom of the girder
yi = vertical distance from soffit to prestress strand layer i
yp = vertical distance from bottom of girder to centroid of nf strands in outer portion
of bulb
a = over capacity factor = Qn/Qu; horizontal tie force fraction; angle of inclination
of transverse reinforcement to longitudinal reinforcement (degrees)
b = factor relating effect of longitudinal strain on the shear capacity of concrete, as
indicated by the ability of diagonally cracked concrete to transmit tension
e = strain (in./in.)
ec = concrete longitudinal strain
ef = failure strain in mild reinforcement (in./in.)
es = net longitudinal tensile strain at the centroid of the longitudinal reinforce-
ment (in./in.)
esh = strain in mild reinforcement corresponding to the beginning of strain harden-
ing (in./in.)
eu = ultimate strain in mild reinforcement (in./in.)
ey = yield strain in mild reinforcement (in./in.)
q = angle of inclination of diagonal compressive stresses (degree); skew angle (degree)
k = multiplier for strand development
ϕ = resistance factor; diameter (in.)
ff = flexure strength reduction factor
fv = shear strength reduction factor
r = reinforcement ratio = Aps /Ag
n = shear stress
rc = unit weight of concrete

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Chapter 1

Background

1.1 Introduction
The initial prestressing force in pretensioned girders can produce large extreme-fiber stresses
particularly near the ends of the girders where span effects (primarily dead load moment) are mini-
mal. The transfer of large prestressing forces at girder ends can additionally lead to local cracking
associated with bursting stresses [transverse stresses due to both Poisson effects in the concrete
and Hoyer effects of the prestressing steel (Briere et al. 2013)] or splitting associated with trans-
fer of the strand force through bond. These effects are compounded by the fact that prestressing
force is introduced to the concrete at a very early age. In many cases, cracking can be mitigated by
permitting greater concrete strength gain prior to prestressing transfer, but this is impractical and
uneconomical within the constraints of the industry and will not be considered further.
AASHTO LRFD Bridge Design Specifications (2010) Article 5.9.4.1.2 provides limits for the
extreme-fiber concrete tensile stress at prestress release. The magnitude of these stresses can be
reduced and associated cracking mitigated in four primary ways: (1) partial debonding (also known
as blanketing or jacketing) a number of strands near the beam end, (2) harping some strands,
(3) adding top strands, and/or (4) a combination of the three preceding ways. Strands that can be
harped are limited to those aligned with the member web(s), which may not be enough to suf-
ficiently lower the stresses. In some cases, such as with boxes, harping is not practical. Harped
strands in relatively thin webs remain susceptible to splitting along their transfer length. More-
over, harping of 0.6-in. or 0.7-in. diameter strands poses challenges in terms of the capacity of
hold-down devices and safety concerns, which are primary reasons some fabricators cite for not
harping strands. The addition of top strands has associated costs and affects the overall stress state
of the section, specifically by counteracting the effects of bottom strands in the critical midspan
region. Top strands can be debonded over the midspan region and cut at midspan after transfer.
This practice, however, complicates production. Partially debonding strands is easily accomplished
using sheathing (flexible split-sheathing or rigid preformed tubes) or greasing, although greasing
may not be entirely effective. The prestressing force in partially debonded strands is transferred
to the concrete at a distance from the end regions where bond is established. In this manner, the
total prestress force is introduced to the member gradually, reducing the stress concentrations and
associated cracking at the beam ends.
Partial debonding of strands decreases the capacity of the longitudinal reinforcement par-
ticularly when cracks pass through the transfer length of debonded strands (Barnes et al. 1999).
Although intended to mitigate various beam end effects, excessive debonding can have detri-
mental effects of its own including the reduction of member flexure and shear capacity as well
as cracking associated with shear or transverse load spreading. In particular, AASHTO LRFD
Article 5.8.3.5 provides requirements for longitudinal steel capacity to resist the combination of
flexure and shear effects as described in Section 1.3.1 of this report.

1  

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

2   Strand Debonding for Pretensioned Girders

The AASHTO LRFD Bridge Design Specifications provides requirements for strand transfer
and development length calculations in Article 5.11.4.2 and for partially debonded strands in
Article 5.11.4.3. At any section, the number of debonded strands should be limited to 25% of
the total number of strands (note that the AASHTO wording is non-mandatory in this instance:
“should,” not “shall”). The commentary notes that a larger percentage based on “successful past
practice” may be considered. For example, Texas permits up to 75% debonding, while North
Carolina allows up to 30%. Arizona, on the other hand, does not allow any partial debonding in
I girders. Currently, there is no consensus regarding the permitted level of debonding. Further-
more, no universally accepted guidelines are available for establishing the layout of debonded
strands, release pattern of the bonded and partially debonded strands, the length of the debonded
regions, or the staggering of the lengths of debonded strands, among other issues.

1.2  Objectives of Research Program


The main objective of the reported study was to develop a unified approach to the design of par-
tially debonded strand regions in prestressed highway bridge girders that would address all aspects of
the service and strength performance of the girder. A multiple component parametric, analytical, and
experimental study was carried out with the objective of demonstrating the effects of strand debond-
ing on prestressed girder design and behavior. Dominant in the performance is the effect that partial
debonding has on the strength (splitting and/or pullout failures), and the shear capacity of the mem-
ber (considering the increased forces carried by the longitudinal strands in the presence of shear).
Emphasis was placed on detailing requirements for partial debonding including, but not limited to,
1) permitted amounts of partial debonding, 2) distribution of partially debonded strands in the cross
section, 3) debonded lengths, 4) locations and staggering of termination of debonded regions, 5) con-
finement of debonded regions and terminations, and 6) influence of the addition of nonprestressed
reinforcement in the region of partial debonding (used to mitigate the effect of partial debonding).
These aspects were investigated through a coordinated analytical and experimental program.
An initial analysis of present AASHTO limits to strand debonding based on previously reported
experimental programs was conducted and followed by a parametric design study, the results of
which informed the decisions and selections made as the analytic and experimental programs pro-
gressed. An extensive finite element method (FEM) modeling program was developed with the goal
of establishing an analytical platform from which a deeper understanding of the local and global
behaviors of prestressed girders containing partially debonded strands could be realized. The result-
ing 3D FEM platform significantly extended the parametric scope of the experimental study. The
effects of strand debonding on the performance of prestressed girders were examined experimen-
tally through design, fabrication, and the testing of both ends of six girders. The FEM platform and a
strut-and-tie model (STM) were utilized to develop a better understanding of the experimental data.
In the process of executing the analytical research, a number of additional questions arose,
including those associated with the transverse behavior of the bulb (irrespective of strand
debonding) and the behavior of highly skewed girders having partial debonding. These issues
were addressed within the overall scope of the project.

1.3  Review of State of the Art and Practice


1.3.1 Longitudinal Reinforcement Requirements Associated
with Shear Capacity
The requirements of Article 5.8.3.5 of the AASHTO LRFD Bridge Design Specifications are central
to the approach taken and results of this study. In the interest of clarity, these are described here.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Background  3  

1.3.1.1  AASHTO LRFD Article 5.8.3.5: Longitudinal Reinforcement


The presence of shear in a section results in a tensile force generated in the primary longi-
tudinal reinforcement that is in addition to that caused by flexure. The free-body diagram of a
diagonally cracked member is conceptually depicted in Figure 1.1.
By summing the moments about point 0 (i.e., the compressive resultant C), the additional
force in the longitudinal reinforcement (T) can be determined (Eq. 1.1).

Vu  Vu 
∑ M = 0; Tdv + Vs ( 0.5dv cot θ ) −
φv
− Vp dv cot θ = 0; ∴T = 
 φv
− Vp − 0.5Vs  cot θ

about C
Eq.1.1

The tensile force T becomes larger as angle q, the inclination of the diagonal crack, becomes
smaller. Including the effects of flexure (away from the supports) and axial load, and assum-
ing that one-half of the applied factored axial load is resisted by the longitudinal reinforcement
(implied in Article 5.8.3.5), the total tensile force is (Eq. 1.2):

Mu N u  Vu 
T= + 0.5 + − Vp − 0.5Vs  cot θ Eq.1.2
dv φ f φc  φv 

Adequate longitudinal reinforcement (summing the contribution of prestressing strands and


nonprestressed reinforcement) must be provided to resist this tension force (Eq. 1.3), i.e.,

Mu N u  Vu 
Aps f ps + As f y ≥ + 0.5 + − Vp − 0.5Vs  cot θ [ AASHTO LRFD Eq. 5.8.3.5-1]
dv φ f φc  φ v 
Eq.1.3

For the end regions of simply supported beams between the inside face of the support to the
critical section for shear, Eq. 1.3 is simplified to (Eq. 1.4):

 Vu 
Aps f ps + As f y ≥  − Vp − 0.5Vs  cot θ [ AASHTO LRFD Eq. 5.8.3.5-2] Eq.1.4
 φv 

Equations 1.3 and 1.4 clearly demonstrate that longitudinal reinforcement is necessary to resist
shear forces. A similar observation can be made from various truss analogy models where the
longitudinal reinforcement is necessary to resist the force from the diagonal compressive struts.

Figure 1.1.   Free-body diagram of forces


after formation of diagonal cracks.
Source: AASHTO LRFD Article C5.8.3.5.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

4   Strand Debonding for Pretensioned Girders

Equation 1.4 must account for the possible lack of full development of the strand and/or
reinforcing steel at the critical section. Additionally, only bonded prestressing strand contrib-
utes to the required longitudinal capacity; thus, greater debonding will require additional non­
prestressed reinforcement as described by Eq. 1.4.

1.3.2  Extant Experimental Studies


A number of previously reported studies have focused on mitigating end region cracking
in prestressed concrete elements (e.g., Ghosh and Fintel 1986; Tadros et al. 2010; Oliva and
Okumus 2010). These studies consider partial debonding as one method of reducing concrete
stresses to acceptable limits. Such studies are relevant in the present context in so far as they
justify the use of partial debonding. Since the premise of this research is that partial debonding
has been found to be necessary, summarizing these studies is not considered necessary.
A relatively large number of past studies have examined shear capacity of prestressed gird-
ers, in particular the influence of debonding on shear capacity. These studies are summarized
as follows:
Kaar, P. H., and Magura, D. D. (1965)—As part of a study focused on evaluating the effects of
partial debonding on performance of prestressed I girders, five specimens having 0%, 33%, and
50% of strands partially debonded were tested to examine their flexural and shear capacities. Two
of the specimens, having 0% and 33% partial debonding, were tested with reduced shear reinforce-
ment to assess the effects of partial debonding on shear capacity. A longitudinal shear failure along
the web-flange interface occurred in the 0%-debonded girder, and a flexure-induced rupture of all
strands at midspan was observed in the partially debonded case. Although the shear capacity was
reported to be “nearly exhausted,” this latter flexural failure was interpreted to indicate that there
was “no detrimental effects of blanketing [debonding] on shear strength.” Strand slip under static
load also led to the conclusion that the development length of partially debonded strands should
be doubled (this latter finding has resulted in k = 2 in AASHTO LRFD Eq. 5.11.4.2.1).
Krishnamurthy, D. (1971)—The effects of strand debonding on the shear strength of preten-
sioned I girders were evaluated. The level of debonding was 25% and 50% per row, and 12.5%,
25%, 37.5%, and 50% per section. The specimens failed suddenly after formation of a diagonal
crack that extended from the support to the load point indicating an abrupt loss of shear capac-
ity. The small-scale specimens contained no shear reinforcement and, therefore, the observed
failure is not believed to be associated with the partial debonding of strands.
Rabbat B. G., Kaar P. H., Russell, H. G., and Bruce, R. N., Jr. (1979)—Fatigue testing of speci-
mens with 10% partially debonded strands indicated slip (attributed to bond fatigue) for the
specimen designed with a development length of the partially debonded strands of d. A similar
specimen designed using a value of 2 d experienced negligible slip. This study confirmed the
appropriateness of the k = 2.0 factor used in AASHTO LRFD Eq. 5.11.4.2.1 and raised the issue
of performance of partially debonded strands in service load (fatigue) conditions.
Csagoly, P. (1991)—Sixteen AASHTO Type IV girders were tested to examine effects of
(1) partial debonding, (2) confinement reinforcing steel in the end zone, and (3) coating of web
reinforcement. The level of partial debonding was 50%. Observed failures were always preceded
by strand slip indicating inadequate anchorage of too little remaining bonded strand.
Abdalla, O. A., Ramirez, J. A., and Lee, R. H. (1993)—The flexural and shear behaviors of pre-
tensioned girders with fully and partially debonded strands were evaluated by testing five AAS-
HTO Type I girders. The test specimens had 0.5-in. diameter strands, and the girder concrete
strength was 6 ksi. The percentage of partially debonded strands was 50% or 67%. All specimens

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Background  5  

were tested over a very short shear span, and failure, in all cases, was characterized by insufficient
anchorage of the prestressing strand regardless of the degree of debonding.

Russell B. W., and Burns N. H. (1993)—The focus of this study was on the transfer and devel-
opment lengths of 0.5-in. and 0.6-in. diameter strands, and to develop guidelines for debonded
strands. Debonding of strands was 50% or less. As expected, specimens with staggered debond-
ing terminations performed better than those with concurrent terminations. It was recom-
mended that no more than 33% of strands be debonded, and at least 6% of the total prestressing
force should be provided in the top flange. No data regarding shear capacity were provided.

Shahawy M., Robinson B., and Batchelor, B. de V. (1993)—As one of its objectives, this study
evaluated the effects of partially debonded strands on shear strength. Thirty-three AASHTO
Type II girders made with 6-ksi concrete were tested. The specimens had three levels of debond-
ing: 0%, 25%, and 50%. The shear span-to-depth ratios (a/d) ranged from 1.3 to 4.8. Four
specimens with 0.6-in. diameter strands having 25% and 50% debonded strands experienced
shear and bond failure. The nominal shear reinforcement in these specimens was equal to that
required by the 1992 AASHTO Standard Specifications for Highway Bridges. The small values of
a/d raise concerns because compression struts forming between the load and the supports domi-
nate the load-carrying mechanism. This research is cited in the commentary to AASHTO LRFD
Article 5.11.4.3 in reference to limiting the total number of partially debonded strands to 25%.

Barnes, R., Burns, N., and Kreger, M. (1999)—The anchorage of 0.6-in. diameter strands was
evaluated by testing 36 AASHTO Type I girders. In addition to concrete strength and strand
surface condition, strand-debonding layout was one of the variables in this study. The partially
debonded strands corresponded to 50%, 60%, and 75% of the total strands. As long as crack-
ing was prevented for a distance of 20 strand diameters beyond the end of the transfer length,
partially debonded strands were observed to resist bond slip. All the specimens failed in flexure,
which was attributed to the excess shear reinforcement preventing loss of bond. Moreover, the
presence of horizontal web reinforcement did not significantly affect the girder behavior because
these bars did not yield by the time the specimens failed in flexure.

Ma, Z., Tadros, M. K., and Baishya, M. (2000)—This research focused on evaluating an accept-
able shear stress limit by performing five shear tests on two NU-1100 I girders with 7.5-in. thick
slabs. The girders used 8-ksi compressive strength concrete and 0.5-in. diameter strands. The
strands were bent up at their ends following release and were cast into a diaphragm. For both
girders, the strands were harped on one end while 13% or 32% of strands were partially debonded
at the other end. Three load points were used to load the girders, with the closest one being 4.5 ft
from the support, resulting in a minimum shear span-to-depth ratio of approximately 1.06. The
load transfer to the support was most likely through arch action in this case. In the first specimen,
the mode of failure for both of the beam ends having harped and partially debonded strands was
web crushing. For the second specimen, the end with harped strands failed due to web crushing.
However, the partially debonded end was stronger to the point that it exceeded the capacity of
the test apparatus. The web reinforcement at the partially debonded end was orthogonal welded
wire fabric. The lack of failure of the debonded ends is attributed to the additional anchorage
provided by the strands (both bonded and debonded) being anchored into the diaphragms. For
simple spans, the end diaphragms for anchoring the strand would likely not be present beyond
the end of the beam.

Nagle, T. J. and Kuchma, D. A. (2007)—As part of NCHRP Project 12-56 (Hawkins and Kuchma
2007), 20 tests were performed on ten 63-in. deep prestressed bulb-tee bridge girders subjected
to uniformly distributed loads. Measured concrete compressive strengths ranged from 9.6 ksi to
16.3 ksi. The average of the test specimens’ calculated capacities was 1.11 times the LRFD-predicted

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

6   Strand Debonding for Pretensioned Girders

capacity. In one specimen, fully bonded straight strands were used at one end, and 38% of the
straight strands were partially debonded at the other end. The end with fully bonded strands
failed at an applied load of 38 k/ft (1.3 times the capacity based on LRFD); additionally, no strand
slip prior to failure was observed. The other end with partially debonded strands failed at
28 k/ft (1.14 times the LRFD capacity); 0.5 in. of strand slip prior to failure was observed. In
another specimen, the same number of strands was used at both ends; however, straight and harped
strands were used at one end, and 23% of the straight strands were partially debonded at the other
end. The end with harped and fully bonded straight strands failed at 43 k/ft (1.34 times the LRFD
capacity) with no strand slip prior to failure. The other end with partially debonded strands failed at
34 k/ft (1.17 times the LRFD capacity) with 0.01 in. of strand slip prior to failure. In this study, the
partial debonding clearly affected beam performance. Note that the specimen with 38% partially
debonded strands had been detailed such that its calculated capacity was 1.01 times the LRFD
capacity, but the failure load exceeded this calculated capacity by 14%.
Moore, A. M. (2010)—This research focused on improving the design and detailing of skewed
end-blocks in Texas DOT prestressed concrete U-beams. One of the objectives was to under-
stand the effect of partially debonded strands on shear performance. Five 54-in. deep U-beams
were tested. In one of the specimens, 46% of the strands were partially debonded (Texas permits
up to 75%). This 30-ft long specimen had no skew at either end and was tested in conjunction
with two girders with fully bonded strands. Irrespective of whether the strands were fully bonded
or partially debonded, the web-to-flange interface failed. As a result, the code-calculated shear
capacity could not be developed. New details were developed to strengthen the web-to-flange
interface, and two additional specimens having only fully bonded strands were tested in the sec-
ond phase of this project. As a result, it was not possible to evaluate the effect of partial debond-
ing on the performance and shear capacity of U-beams.
Morcous, G., Hanna, K., and Tadros, M. K. (2010)—The focus of this study was on the influ-
ence of confinement reinforcement on the transfer length, development length, and shear capac-
ity of prestressed bridge girders. Approximately 25% of the strands were partially debonded at
one end. The amount and distribution of confinement reinforcement was found to have an
insignificant effect on the transfer and development lengths, or the ultimate flexural or shear
resistance, beyond the shear contribution of the confining reinforcement.
Tadros, M. K., and Morcous, G. (2011)—This study examined a number of issues related to
the use of 0.7-in. diameter strands in pretensioned girders. At one end, 25% of the strands were
partially debonded. It was concluded the shear capacity could conservatively be predicted using
the AASHTO LRFD Bridge Design Specifications if the following conditions are satisfied: (1) the
concrete has a minimum compressive strength of 10 ksi, (2) the number and pattern of partially
debonded strands are in accordance with Article 5.11.4.3, and (3) the provided bottom flange
reinforcement is at least equal to that prescribed in Article 5.10.10.2.
Wesson, M. (2013)—A project at Purdue University focused on understanding the effects of
debonding on the nominal shear strength provided by concrete when diagonal cracking results
from combined shear and moment (Vci) and the nominal shear strength provided by concrete
when diagonal cracking results from high principal tensile stress in web (Vcw). The experimental
program consisted of four phases. Four beams were tested for each phase, having debonding
ratios of 0%, 25%, 50%, and 75%. In Phase I, the rectangular beam test specimens were designed
such that location of critical Vci was located within the region with partially debonded strands. A
significant reduction of capacity was observed when 75% of the strands were partially debonded.
In Phase II, rectangular test beams were designed so that the location of critical Vci would be
located outside of the region where the strands were partially debonded. The beams with 25%
and 50% debonding performed similarly to the beam with 0% debonding. The beam with 75%
debonding showed a significant reduction in capacity. I girders were tested in Phase III to evaluate

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Background  7  

the effects of partial debonding on Vcw. Again the beams with 25% and 50% debonding performed
similarly to the beam with 0% debonding and the beam with 75% debonding showed a significant
reduction in capacity. In Phase IV a single rectangular beam with 50% debonding was tested to
evaluate the impact of having transverse reinforcement in the expected failure region. In the test
specimens, the debonded strands were not staggered. None of the specimens met the longitudinal
reinforcement requirements of AASHTO LRFD Article 5.8.3.5.

1.3.2.1  Synthesis of Past Experimental Studies


Past studies involving partially debonded strands lack a clear consensus of the effects and
limits of partial debonding practice. However, a number of trends are clearly evident.
1. The effect of partial debonding on the development of longitudinal flexural reinforcement
is critical (e.g., Kaar and Magura 1965; Rabbat et al. 1979; Csagoly 1991; Russell and Burns
1993; Barnes et al. 1999). This observation implies that the span characteristics such as the
moment-to-shear ratio play a significant role in member behavior. Strand slip and splitting
failures are indicative of this response. The observation of strand slip raises issues associated
with service load behavior in addition to strength capacity.
2. Partial debonding has little or no effect on the methodology for the calculation of the shear
resistance of the member (e.g., Kaar and Magura 1965; Abdalla et al. 1993; Barnes et al. 1999;
Nagle and Kuchma 2007). That is, the combination of Vc and Vs remains unaffected. Although
shear is recognized to increase the stress in the longitudinal strand, this effect leads back to
the need to fully develop the bonded strand as described in point 1.
One reason that partial debonding may have little effect on shear capacity is that near the
debonded end region, shear behavior becomes dominated by arching or direct strut action with
the anchorage provided at the beam support itself (e.g., Shahawy et al. 1993; Ma et al. 2000).

1.3.3  Past Analytical Studies


Modeling of global prestressed girder behavior is commonplace using finite element
approaches. Baxi (2005) developed numerical models using ABAQUS to study the bond behavior
in prestressed concrete members with fully bonded or partially debonded strands. Baxi imple-
mented a concrete model based on the commercial software ANACAP-U (using a user material
subroutine in ABAQUS) without considering the concrete creep and shrinkage effects. The pre-
stressing strand was modeled using truss elements, and its bond with the concrete approximated
using discrete spring elements. A nonlinear law was used to describe the shear resistance of the
spring element.
Subsequently, Burgueño and Sun (2011) improved the numerical models substantially by
accounting for creep and shrinkage, and by replacing spring elements with a more realistic contact
surface to model bond. In their study, the damage-plasticity model provided by ABAQUS was
used for concrete. Prestressing strands were modeled using 3D solid elements, and their bond with
concrete was approximated using a contact surface. The basic Coulomb friction model was used to
capture the Hoyer effect. The creep and shrinkage strains were not obtained from FEM simulation
or experimental tests, but rather were based on simple calculations of the CEB-FIP (1990) model.
This approach was very successful for analysis of small-scale beams. However, modeling of strands
with 3D solid elements requires mesh sizes that are computationally unaffordable for analysis of
full-scale girders. To overcome this obstacle, Burgueño and Sun had to limit the use of 3D solid
elements for concrete and strands in a small part of the girder while the major part of the girder
was meshed using shell and truss elements. Unfortunately, shell elements are a poor selection to
model concrete fracture, and, thus, lead to inaccurate prediction of the overall behavior of the
girder under complicated loading scenarios.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

8   Strand Debonding for Pretensioned Girders

1.3.3.1  Synthesis of Past Analytical Studies


Past studies have not adequately addressed issues associated with modeling of bond and local
effects in the vicinity of strands. Furthermore, neither 3D nonlinear finite element models nor
simpler techniques, such as modified compression field theory or STMs, have been used to
investigate the effects of partial debonding on shear capacity.

1.4  Current Practice for Debonding Strand


AASHTO LRFD Article 5.11.4.3 provides the following requirements for partially debonded
strands:
• The development length, measured from the end of the debonded zone, shall be determined
using Eq. 5.11.4.2-1 with a value of k = 2.0 (italics added to emphasize the wording).
• The number of partially debonded strands should not exceed 25% of the total number of
strands.
• The number of debonded strands in any horizontal row shall not exceed 40% of the strands
in that row.
• Not more than 40% of the debonded strands or four strands, whichever is greater, shall have
their debonding terminated at any section.
• The length of debonding of any strand shall be such that all limit states are satisfied with
consideration of the total developed resistance at any section being investigated.

1.4.1  State Amended Specifications


The design manuals from a number of states were reviewed. The following brief summary
indicates a wide range of requirements for partially debonded strands:
Arizona (2011)—Partial debonding shall not be allowed for I girders.
Indiana (2011)—Different levels of partial debonding are allowed. For bulb-tee beams, the
limit is 25%; however, for AASHTO I girders and box beams, as much as 50% of the strands can
be partially debonded.
Nebraska (2011)—An increase of less than 10% beyond AASHTO limits of 25% and 40% is
allowed with the permission of the Assistant Bridge Engineer.
New York (April 2006)—The requirements of AASHTO LRFD Article 5.11.4.3 are specified
with the following additions:
• Spacing of partially debonded strands shall be a minimum of 100 mm (4 in.).
• At a given section, it is permitted to rebond a maximum of four partially debonded prestress-
ing strands. A minimum difference of 600 mm (24 in.) is required between debonding lengths.
• Partial debonding is not permitted in units 380 mm (15 in.) or less in depth.

North Carolina (October 2011)—AASHTO LRFD Article 5.11.4.3 is specified with the fol-
lowing relaxation of the 25% limit: “The number of debonded strands shall preferably not
exceed 25% but never more than 30% of the total number of strands.”
Ohio (July 2007)—AASHTO LRFD Article 5.11.4.3 is specified. However, the 25% limit on
the number of debonded strands is mandatory for I-beams (in contrast to AASHTO, which uses
should rather than shall). In addition, the following details are specified:
• The maximum debonded length at each beam end shall not be greater than 0.16L - 40 in.,
where L equals the span length in inches.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Background  9  

• A minimum of one-half the number of debonded strands shall have a debonded length equal
to one-half times the maximum debonded length.
• Strands extended from a beam to develop positive moment resistance at pier locations shall
not be debonded strands.
Pennsylvania (2000)—Partially debonded strands in lieu of draping is limited to 25% of the
total number of strands. Pennsylvania Design Manual 4 describes “crack control debonding”
which is limited to 50% of the total strands within 6 in. of the end of the beam and 25% through
the first 36 in. This crack control debonding is in addition to the 25% limit of Article 5.11.4.3
although no more than 50% of strands may be debonded at any section.
Texas (2010)—AASHTO LRFD Article 5.11.4.3 is modified to permit up to 75% partial
debonding and as many as 75% of strands in a given row. Debonding termination must be
staggered such that no more than 75% of the debonded strands or 10 strands are “rebonded”
at any section. Additionally, the maximum debonding length is limited to the lesser of one-half
the span length minus the maximum development length, 20% of the beam length, or 15 ft; and
strands are debonded in 3-ft increments.

1.4.2  Survey of Current Practice


A detailed survey was developed and distributed to all state DOTs, two Canadian provincial
ministries of transport (MOTs) (Alberta and Ontario), and South Korea. The survey instru-
ment was modeled after a survey from the Precast/Prestressed Concrete Institute (PCI) and
was designed to compile a summary of current practice for partially debonded strands. The
survey instrument is provided in Appendix A. The response rate was 66% with 35 state DOTs
responding. Based on the results summarized in Appendix A, the following primary observa-
tions are drawn:
a. 77% of the states (27 states) that responded allow partial debonding; 17% (6 states) do not
permit partial debonding; while 6% (2 states) are neutral on the subject. One state allowed
partial debonding in the past although no explanation was provided for why it is no longer
permitted.
b. The most common technique for partially debonding of strands is split plastic sheath-
ing (54%—19 states) followed by preformed plastic tubes. None of the respondents use
debonding agents.
c. Most states (54%—19 states) have no preference between partial debonding or harping. Of
those with a preference, 13 states (37%) prefer harping; partial debonding is preferred by
only three states (9%).
d. When harping and partial debonding options are both available, 41% (14 states) prefer harp-
ing in comparison to 20% (7 states) for partial debonding. 14% (5 states) of the respondents
indicate that there is no consistent local practice for either option.
e. The most common diameter of strands used is 0.6-in. (93%—33 states) followed by 0.5-in.
diameter strands (85%—30 states). None of the states reporting regularly permit the use of
0.7-in. diameter strands except for 3 states that allow their use by special permission. The
majority of the respondents (89%) do not use 0.375-in. diameter strands.
f. The following is the number of respondents using various girder shapes:
i. Bulb-Tee Girders 30 (86%)
ii. I Girders 28 (79%)
iii. Box Girders/Voided Slabs 27 (76%)
iv. Tee Girders (includes single or double tee, deck bulb-tee, or tri beams) 12 (34%)
v. U- (or trapezoidal) Girders   8 (24%)

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

10   Strand Debonding for Pretensioned Girders

g. The number of states using different permitted percentages of partial debonding is as follows:

(a) % of all strands in girder


Shape 0% 25%1 30% 33% 40% 50% 75%
I Girders 2 9 1 2 1 3 0
Bulb-Tees 2 14 1 2 1 3 0
U- (or Trapezoidal) Girders 0 4 0 2 0 0 1
Box Girders/Voided Slabs 0 12 1 2 1 1 1
Tee Girders 1 7 0 0 0 0 0
(b) % of strands in an individual row of strands
Shape 0% 25% 30% 33% 40%1 50% 75%
I Girders 2 1 0 0 10 4 0
Bulb-Tees 2 1 15 0 0 4 0
U- (or Trapezoidal) Girders 0 1 0 0 4 1 1
Box Girders/Voided Slabs 0 1 0 0 13 3 1
Tee Girders 1 0 0 0 7 1 0
1
AASHTO-prescribed limit.

Although the majority of states use the AASHTO limits of 25% total and 40% in a single
row, some states are in the process of increasing these limits. As of August 2016, Florida
allows a maximum percentage of partially debonded strands of 30%, with commentary that
this is a conservative interim limit pending results from this project. In these cases, the cur-
rent AASHTO limits are deemed to be unnecessary as long as longitudinal reinforcement per
AASHTO LRFD Article 5.8.3.5 is provided.
h. Although there are preferences in terms of girder shapes (f, above), no responding state
appears to differentiate debonding limits based on girder shape, or indeed on any other fac-
tor except the total strands in a section and the number of strands in a row as described in
AASHTO LRFD Article 5.11.4.3.
i. The majority of respondents do not report any cracking during fabrication or while the
girder is in service with the exception of a few rare cases (5% of girders or fewer per year).
Cracking does not appear to be a main concern. For instance, the reply from Massachusetts
states that, “Inspectors would not be expected to attribute cracking to debonding. In fact, end
girder cracking is very common.”

No. of states reporting cracking in fewer than


Shape 5% of girders
During fabrication In service
I Girders 4 1
Bulb-Tees 1 1
U- (or Trapezoidal) Girders 1 1
Box Girders/Voided Slabs 1 1
Tee Girders 2 2

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Chapter 2

Analytical Research Approach


and Findings

2.1  Research Approach


A multiple component analytical study was carried out with the objective of demonstrating
the effects of strand debonding on prestressed girder design and behavior.
An initial analysis of present AASHTO limits to strand debonding based on two previously
reported experimental programs was conducted (Section 2.2). Following this analysis, a parametric
design study (Section 2.3) was undertaken in which preliminary designs of 1,392 girders were
performed, of which 522 cases were selected for more detailed evaluation based on AASHTO
LRFD Bridge Design Specifications provisions. The parameters considered in this study included
(1) girder shape, (2) strand diameter, and (3) concrete strength. For each case, four amounts
of total prestressing were considered: 100% of strands that may be located in the bulb or lower
flange (Nmax) and approximately 75%, 50%, and 25% of this value. The results of this parametric
study provided information for the decisions and selections made as the analytic and experimental
programs progressed.
An extensive FEM modeling program (Section 2.4) was developed with the goal of developing
an analytical platform from which a deeper understanding of the local and global behaviors
of prestressed girders containing partially debonded strands could be established. The result-
ing 3D FEM platform significantly extended the parametric scope of the experimental study.
Forty-four prototype girders (and some additional variations on these) were modeled, and their
performance was evaluated. The FEM platform was validated using experimental data available
in the literature, and its efficacy was further confirmed by accurately modeling the experimental
behaviors observed in this study (Chapter 3).
In the process of executing the analytical research, a number of additional questions arose.
Issues associated with the transverse behavior of the bulb were addressed through an extensive
STM-based study (Section 2.5), which included 9,408 design cases. The results of the STM study
were used to refine detailing recommendations for prestressed concrete girders irrespective of
strand debonding (Chapter 3).
Finally, a question raised during the research regarding debonding of girders with highly skewed
ends was addressed (Section 2.6) using techniques of fundamental mechanics and validated
using the FEM platform developed for the research program.

2.2 Evaluation of Current AASHTO Limits


on Strand Debonding
The results from two previous studies are used to further understand the influence of strand
debonding. These studies are presented in the following sections.

11  

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

12   Strand Debonding for Pretensioned Girders

2.2.1  Shahawy et al. (1993)


The recommended 25% debonding limit in the current AASHTO LRFD Bridge Design Specifi-
cations is attributed to an experimental study by Shahawy et al. (1993). In this study, 32 full-scale
prestressed girders were tested, including 8 girders with debonded strands: 4 having a debonding
ratio, dr = 0.25 and 4 having dr = 0.50. Both the north and south ends of each girder were tested;
thus, a total of 16 tests examined debonded conditions [although only 14 data points (7 girders)
were reported]. Table 2.1 summarizes the failure loads observed for each test of a girder having
debonded strands and the corresponding “control” girders having no debonding. The data
indicate no clear trend, especially between the girders having dr = 0.25 and those having dr = 0.50;
each failed at very similar loads. Shahawy et al. (1993) based their recommendation for limiting
debonding to 25% on the failure modes of the 50% debonded girders, which were described as
being less ductile.
The Shahawy et al. test girders were designed based on the AASHTO Standard Specifications for
Highway Bridges (1989), which did not require a check for longitudinal tension as is now required by
AASHTO LRFD Article 5.8.3.5. After publication of the results of their tests for the fully bonded
girders [Shahawy and Batchelor (1996)], Collins and Mitchell (1997) pointed out in a discussion
that the girders did not meet the longitudinal reinforcement requirements implied by the modified
compression field theory; these requirements were later adopted as AASHTO LRFD Article 5.8.3.5.
A review of Shahawy et al. (1993) confirmed that the debonded girders did not meet the longitu-
dinal reinforcement requirement of AASHTO LRFD Article 5.8.3.5. Thus, it is possible that the
undesirable failure mode reported was the result of failing to meet the longitudinal tension steel
requirements, which would most likely be accentuated by debonding strands in some girders.
It is believed that the failure mode may have been different had additional, nonprestressed longi-
tudinal reinforcement, as required by the current AASHTO LRFD Specifications, been provided.
The effects of the specimens not meeting the longitudinal reinforcement requirements were
evaluated, as follows:
1. The girder parameters (area of bonded prestressing steel; concrete strength; girder depth, h;
shear depth, dv; etc.) were determined from Shahawy et al. (1993). The distance to the critical
section, dc, was taken as dv from the center of bearing consistent with the 2016 version of the
AASHTO LRFD Bridge Design Specifications.
2. The total shear (Vtotal) and moment (Mtotal) at the critical section, dc, were determined as the sum
of the effects of the experimentally applied concentrated load (P) and the girder self-weight.
3. Using Vtotal and Mtotal at dc, the corresponding values of es, q, and b were calculated based on
the current version of AASHTO LRFD Article 5.8.3.4.2. These values were used to calculate Vc.
4. The total shear resistance, Vn = Vc + Vs, was found and compared to the value of Vtotal. The
value of P was iterated upon (Steps 2 and 3) until Vtotal = Vn. This value is the predicted shear
capacity of the girder, Vn.
5. The longitudinal tension at the critical section was checked according to AASHTO LRFD
M total  Vtotal 
Eq. 5.8.3.5-1, i.e., Aps f ps ≥ + − 0.5Vs  cot q (in these tests Nu and Vp were zero).
dv φ f  φv 
The value of P was again iterated upon (and es, q, and b changed accordingly since q affects Vs)

Table 2.1.   Failure load (kips) for Shahawy et al. (1993) tests.
dr = 0 dr = 0.25 dr = 0.50
Test Series
North end South end North end South end North end South end
A0-XX-R 313 276 281 173 not reported
A2-XX-3R 257 312 253 170 200 160
C0-XX-R 176 180 237 123 233 127
C1-XX-R 177 196 164 143 158 133

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Analytical Research Approach and Findings   13  

until the two sides of the equation are made equal. This new value of Vtotal is the predicted
girder capacity if the failure is due to insufficient longitudinal steel at the critical section, VT@dc.
6. Finally, the longitudinal tension at the face of the support was checked according to AASHTO
 Vtotal 
LRFD Eq. 5.8.3.5-2, i.e., Aps f ps ≥  − 0.5Vs  cot q. Once again, the value of P was iterated
 φv 
upon until the two sides of the equation were made equal. This final value of Vtotal is the
predicted girder capacity if the failure is due to insufficient longitudinal steel at the face of the
support, VT@support.
The three predicted values of shear to cause failure (Vshear, VT@dc, and VT@support) were compared
to the experimentally reported shear values at failure, Vexp, at each section considered (dc and face
of the support) for each girder reported by Shahawy et al. (1993). Figure 2.1 shows plots of
the calculated and observed capacities for all three scenarios. Data falling “above” the 1:1 line
indicate that the observed failure shear was lower than the calculated shear capacity; values falling
below the 1:1 line indicate that the observed capacity exceeded the calculated capacity. Data for
all girders for which data were available are shown; girders having debonded strands are shown
with solid data points. The following observations are made:
1. Fifty-two of the 64 specimens tested did not achieve the shear strength capacity (Vn) predicted
by AASHTO LRFD Bridge Design Specifications (2016) (Figure 2.1a).
2. The experimentally observed strength of the girders is best predicted by the available longitu-
dinal steel tensile capacity of the girder at the critical section (VT@dc, shown in Figure 2.1b). This
conclusion is consistent with the observations of Collins and Mitchell (1997). These girders
were designed based on the Standard Specifications for which there was no requirement
for minimum longitudinal steel.
3. The predicted capacity based on longitudinal tension at the face of the support (VT@support) is
seen to be very conservative (Figure 2.1c). This observation is not unexpected. This provision

450
calculated capacity
overestimates
400 observed capacity

350
Calculated Shear Capacity (kips)

Calculated
300 = 1.31
Observed

250

200

150 25% debonding


50% debonding

100

50 calculated capacity
underestimates
observed capacity
0
0 50 100 150 200 250 300 350 400 450
Observed Shear Capacity (kips)
(a) shear capacity (Vn) calculated using AASHTO LRFD Eq. 5.8.3.4.2

Figure 2.1.   Experimental and predicted shear capacities of Shahawy


et al. (1993) girders.
(continued on next page)

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

14   Strand Debonding for Pretensioned Girders

450
calculated capacity
overestimates
400 observed capacity

Calculated Shear Capacity based on T @ dcrit (kips)


350

300 25% debonding


50% debonding
Calculated
250 = 0.81
Observed

200

150

100

50 calculated capacity
underestimates
observed capacity
0
0 50 100 150 200 250 300 350 400 450
Observed Shear Capacity (kips)
(b) shear capacity at critical section (VT@dc) calculated using AASHTO LRFD Eq. 5.8.3.5-1

450
calculated capacity
Calculated Shear Capacity based on T @ support face (kips)

overestimates
400 observed capacity
25% debonding
350 50% debonding

300

250

200 Calculated
= 0.46
Observed
girders having
150 '2-point loading'

100

50 calculated capacity
underestimates
observed capacity
0
0 50 100 150 200 250 300 350 400 450
Observed Shear Capacity (kips)
(c) shear capacity at face of support (VT@support) calculated using AASHTO LRFD Eq. 5.8.3.5-2

Figure 2.1.  (Continued).

of the AASHTO LRFD Bridge Design Specifications assumes a linear bond relationship over the
strand transfer length of 60db. This transfer length is known to be conservative resulting in an
under-prediction of the tensile resistance provided in situ.
4. The predicted VT@dc and VT@support capacities (Figures 2.1b and c) of girders having debonded
strands are lower than comparable girders reflecting the reduced Aps term in AASHTO LRFD
Eqs. 5.8.3.5-1 and 5.8.3.5-2 due to the debonded strands. The fact that the experimentally
observed capacity remains essentially unchanged, as shown in Table 2.1, further demonstrates

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Analytical Research Approach and Findings   15  

that the debonding itself had little effect on girder capacity provided that the remaining
bonded steel is adequate to resist the tension demand.
5. The data are inconclusive. It is not possible to conclude that debonding had no effect on the
shear strength; however, it is also not possible to conclude that debonding was the only cause
of premature shear failure.

2.2.2  Russell et al. (2003)


As part of this study, both ends of three 72-in. deep bulb-tee girders were tested. Each beam
had 24 strands with 6 strands debonded at the ends, resulting in 25% debonding. Confinement
reinforcement in the bottom flange at the ends consisted of No. 3 bars at 6 in. spacing. One speci-
men (BT6) was designed using the AASHTO Standard Specifications for Highway Bridges (1996).
As such, no additional longitudinal reinforcement at the ends was included. The other two girders
(BT7 and BT8) were designed using the AASHTO LRFD Bridge Design Specifications through the
2000 Interim Provisions, which required the addition of 8 No. 6 longitudinal bars at the ends.
For specimen BT6, the load capacity was limited by strand slip. One end of girders BT7 and
BT8 could not be tested to failure because the capacity of the test apparatus was reached. Up to the
maximum load that could be applied, BT7 exhibited some slippage whereas BT8 did not. Testing to
failure of the other end indicated no slippage in BT8 and some slippage in BT7. Both girders failed
at the web-to-bottom flange interface with web crushing/spalling and horizontal shear.
Russell et al. (2003) confirmed that if there is not sufficient longitudinal reinforcement, the
strands will slip. This slippage results in a shear-bond failure not a shear failure. A number of
the specimens (with 25% or 50% debonding) tested by Shahawy et al. (1993) also exhibited
shear-bond failures. Shear-bond failures are not deemed to be a rational basis for limiting strand
debonding provided the longitudinal reinforcement is adequate.

2.3  Design Case Studies


In order to better understand the effects of debonding across a range of girder types and
capacities, a parametric design study was undertaken. The parameters considered in the study
(Table 2.2) included (1) girder shape, (2) strand diameter (db), and (3) concrete strength ( f c′).
Composite slab thickness (t) and girder spacing (S) were constant for each girder type. For each
case, four amounts of total prestressing were considered: 100% of strands that may be located in
the bulb or lower flange (Nmax) and approximately 75%, 50%, and 25% of this value.
All analyses were carried out considering the same loading in which permanent loads, DC,
were taken as the girder self-weight, slab weight, and an additional 300 lb/ft to account for
barrier wall loads (two 750 lb/ft barriers distributed over five girders). DW was taken as 35 psf
(3-in. thick asphalt overlay). HL-93 live loading was assumed and the maximum moment
calculated at midspan.1 Distribution factors were calculated for interior girders having two or
more lanes loaded.

2.3.1  Determination of Maximum Girder Span


The intent of this study was to evaluate large debonding requirements with girders being
“stretched” to their maximum span lengths and capacities. The maximum girder span length for
each section was taken as the greatest length satisfying AASHTO LRFD STRENGTH I, SERVICE I,

1 
It is acknowledged that the maximum live load moment does not occur at the midspan; however, this assumption simplifies
the calculations, and is a very close estimate of the maximum live load moment, especially for longer spans.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

16   Strand Debonding for Pretensioned Girders

Table 2.2.   Range of values of parameters considered.


Range of Values
Parameter I girders Bulb-tees Adjacent and U-girders
Spread Boxes
Girder type Type II, IV and BT-54 BI-36, BII-36, T-U40
VI; WF36-49, BT-63 BIV-36 T-U54
WF54-49, WF72- BT-72 BI-48, BII-49, W-U54G4
49, NU-900, NU- BIV-48 W-U66G4
1350, NU-1800 W-U78G4
Total number of strands in section Nmax, 0.75Nmax, 0.50Nmax and 0.25Nmax
S: Girder spacing 8 ft 8 ft 8 ft (spread) 14 ft
4 ft (adjacent)
t: Cast-in-place slab thickness 8 in. 8 in. 8 in. (spread) 8 in.
3 in. (adjacent)
L: Simple span length Minimum practical to the maximum possible value (Section 2.3.1)
dr: Debond ratio (Section 2.3.2) 0 to 0.76 0 to 0.77 0 to 0.77 0 to 0.78
Concrete material properties f’c = compressive strength of concrete = 6, 8, 12, 15 ksi
f’ci = concrete compressive strength at prestress transfer = 0.6f’c
f’c (deck) = compressive strength of composite cast-in-place deck concrete = 5 ksi
c = unit weight of all concrete = 0.15 kcf
Strand db = 0.5-, 0.6-, 0.7-in. diameter low-relaxation strand
fpu = ultimate strength of prestressing strand = 270 ksi and Eps = 28500 ksi
strands are initially stressed to 0.75fpu = 202.5 ksi
fpi = initial prestress at transfer = (0.9)0.75fpu = 182 ksi (10% loss at transfer)
Long-term effective stress in the prestressing strand = 0.56fpu = 151 ksi

and SERVICE III requirements; that is, L = min[LSTRENGTH I , LSERVICE I , LSERVICE III] ≥ Lmin, in which
each length is defined in the following sections. In the following equations, compressive stresses
are taken as positive.

2.3.1.1  STRENGTH I
The design capacity (fMn) of all girder sections was determined using plane section analysis
software RESPONSE 2000 (Bentz 2000). From this, the maximum permissible span length of
the girder (LSTRENGTH I) was determined as the greatest value of span length satisfying fMn/Mu ≥ 1.0
for the STRENGTH I load combination.

2.3.1.2  SERVICE I (AASHTO LRFD Articles 3.4.1 and 5.9.4.2.1)


Under SERVICE I loading, compression stresses are limited to 0.45f c′ for the effects of the
sum of effective prestress and permanent loads and to 0.60f c′ for the effects of the sum of effective
prestress, permanent loads, and transient loads. The compressive stresses were computed as
follows (Eqs. 2.1–2.2):
Girder:

 1 e girder  M girder + M slab MWS + M barrier wall


NAp f pLT  − + + ≤ 0.45 f c′, girder Eq. 2.1a
 Agirder Stop of girder  Stop of girder Stop of girder , composite

 1 e girder  M girder + M slab MWS + M barrier wall + M LL


NAp f pLT  − + + ≤ 0.60 f c′, girder Eq. 2.1b
 Agirder Stop of girder  Stop of girder Stop of girder , composite

Deck slab:

MWS + M barrier wall


≤ 0.45 f c′,slab Eq. 2.2a
Stop of slab , composite

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Analytical Research Approach and Findings   17  

MWS + M barrier wall + M LL


≤ 0.60 f c′,slab Eq. 2.2b
Stop of girder , composite

The subscript “girder” refers to the properties of the precast girder alone whereas the subscript
“composite” refers to the full composite section. From the calculations required for Eqs. 2.1 and 2.2,
a maximum girder span length (LSERVICE I) was determined for the SERVICE I load combination.

2.3.1.3  SERVICE III (AASHTO LRFD Articles 3.4.1 and 5.9.4.2.2)


Under SERVICE III loading, tension stresses in precompressed tension zones having bonded
reinforcement, calculated by Eq. 2.3, are limited to 0.19√f c′.

1 e girder  M DC M DW + M LL
NAp f pLT  +  − − ≥ − 0.19 f c′, girder Eq. 2.3
 Agirder Sbottom of girder  Sbottom of girder Sbottom of girder , composite

Equation 2.3 was used to determine a maximum girder span length (L SERVICE III) for the
SERVICE III load combination.

2.3.1.4  Minimum Girder Span Length


At prestress transfer, the use of Nmax (particularly with 0.7-in. diameter strands) results in
large compressive stresses in the bottom flange. In order to meet AASHTO specified compressive
stress limits, a minimum girder span length (Lmin) is required so that the tensile stresses from the
girder dead load (Mgirder) counteract these compressive stresses. However, the girder span length
is also limited by STRENGTH I, SERVICE I, and SERVICE III requirements. A workable case
cannot be identified if the span length from strength and service requirements is less than that
required to decompress the heavily prestressed flange. This scenario occurred for many of the
cases considered.

2.3.2  Debonding Ratio


Having established the maximum girder span length, L = min [LSTRENGTH I , LSERVICE I , LSERVICE III] ≥
Lmin, the debonding requirement for each case may be determined. The maximum number of
strands (Nt) that may be placed in the section and continue to satisfy the 0.24√f ci′ concrete tensile
stress limit at prestress transfer at the location of the transfer length (Lt) is determined from
(Eq. 2.4):

1 e girder 
N t Ap f pi  − ≥ − 0.24 f ci′ Eq. 2.4
 Agirder Sbottom of girder 

The use of the 0.24√f ci′ limit requires the addition of nonprestressed steel to resist cracking.
Significantly more debonding will be required if the lower tensile limit (requiring no additional
steel) of 0.0948√f ci′ ≤ 0.2 ksi is adopted. Nt is a “property” of the section geometry and concrete
strength at prestress transfer only. The concrete compression limit (Eq. 2.5) at prestress transfer
of 0.6f ci′ was also checked at the transfer length (Lt) and, as expected, found not to control in
any case. The required strand-debonding ratio in each case is, therefore, dr = 1 - Nt /N where
N is the number of strands provided. Required debonding ratios were determined ranged up to
greater than 75%.

1 e girder 
N t Ap f pi  + ≤ 0.60 f ci′ Eq. 2.5
 Agirder Sbottom of girder 

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

18   Strand Debonding for Pretensioned Girders

The concrete compression limit of 0.6f ci′ and tension limit of 0.24√f ci′ were also checked at
midspan at the time of prestress transfer. In this calculation, the total number of strands in the
section, N, was considered and the self-weight of the girder was included. The compressive and
tensile stresses at midspan were determined from Eqs. 2.6 and 2.7, where the moment at midspan,
Mgirder, is that resulting from the self-weight of the girder only (i.e., Mg = Ag rc L2/8; where rc is the
density of concrete).

1 e girder  M girder
NAp f pi  − + ≥ − 0.24 f ci′ Eq. 2.6
 Agirder Stop of girder  Stop of girder

1 e girder  M girder
NAp f pi  + − ≤ 0.60 f ci′ Eq. 2.7
 Agirder Sbottom of girder  Sbottom of girder

For a number of cases with the largest number of strands (Nmax) and lowest concrete strength
( f c′ = 6 ksi), the top tensile stress at the midspan exceeded the AASHTO limit. These cases were
not considered further.

2.3.3  Summary of Design Parameter Study


Initially, 1,392 cases were considered based on combinations of the parameters shown in
Table 2.2. From these, 522 cases (171 I girders, 218 box girders, and 133 U-girders) were identi-
fied for further analyses using a series of MATLAB scripts written specifically for this project.
The complete summary of analysis results is provided in Appendix B.
Figure 2.2 summarizes the reinforcement ratio (r = Aps /Ag; where Aps is the total area of strands
in the sections, bonded and unbonded, and Ag is the girder gross cross-sectional area) of these cases
in terms of debonding ratio (dr) and span length normalized with respect to the girder depth
(L/h). As expected, a larger amount of debonding is required for cases with a large value of r that
correspond to those girders having a larger L/h ratio; i.e., girders for which flexure dominate the
response.
All of the selected cases required partial strand debonding, ranging between 3% and 77%.
(The cases that satisfied the prestress transfer stress limits without partial debonding were not
included in the analyses.) Debonding lengths were chosen to satisfy stress limits, and a 3-ft
debonding increment was used.
Additional nonprestressed reinforcement was added as required to satisfy the longitudinal
reinforcement requirements of AASHTO LRFD Article 5.8.3.5. As a first attempt, No. 4 Gr. 60
reinforcing bars were considered. If No. 4 bars did not provide sufficient increase in the tensile
strength, larger Gr. 60 bars (No. 5 and No. 6) were evaluated. The development lengths were
checked and found to be adequate. It should be noted that it would not be possible to add additional
steel for cases with 100% of strands in the bulb or lower flange, i.e., those with N equal to Nmax.
In Figure 2.3, the ratio of available longitudinal resistance to tensile demand (Aps fps /T) is plotted
against the debonding ratio (dr) and span normalized with respect to the girder depth (L/h). The
following observations are made:
1. As expected, the value of Aps fps /T decreases as the level of debonding ratio (dr) increases.
2. The lowest values of Aps fps /T are for cases with the smaller values of L/h. This trend is also
expected because the influence of shear, which increases the tensile demand (T), becomes
more pronounced as L/h decreases.
3. Cases with the larger debonding ratios (dr) correspond to those with the smaller L/h. For
example, based on the best linear fit through the results of all I girders, dr = 0.25 and 0.70

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Analytical Research Approach and Findings   19  

(a) I girders and bulb-tees

(b) box girders

(c) U-girders

Figure 2.2.   Reinforcement ratio versus debonding ratio and normalized span length.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

20   Strand Debonding for Pretensioned Girders

(a) I girders and bulb-tees

(b) Box girders

(c) U-girders

Figure 2.3.   Variation of Apsfps/T as a function of debonding ratio and normalized span length.
(No nonprestressed reinforcement).
Copyright National Academy of Sciences. All rights reserved.
Strand Debonding for Pretensioned Girders

Analytical Research Approach and Findings   21  

(No nonprestressed reinforcement)

Figure 2.4.   Interrelationship between Apsfps/T, dr, and L/h. (I girders with no nonprestressed
reinforcement shown).

approximately correspond to L/h = 23.2 and 10, respectively (Figure 2.4). This trend is
attributed to the fact that longer girders have larger dead loads that counteract the effects of
the prestressing force; hence, a short, heavily prestressed girder requires a higher degree of
debonding.
4. The largest and smallest values of Aps fps /T are for box and U-girders, respectively.
5. There are no major differences between similar girder types.
6. Relatively few numbers of reinforcing bars were found to be necessary to remedy the tensile
strength deficiency resulting from partially debonded strands. The maximum number of bars
required for each girder type is as follows:
a.  I girders and bulb-tees: 9 No. 4 Gr. 60 for Type IV
(L = 120 ft, f c′ = 15 ksi, dr = 0.72, 0.6-in. diameter stands, N ≈ 0.75Nmax)
b.  Box girders: 5 No. 4 Gr. 60 for BI-36 spread box
(L = 45 ft, f c′ = 6 ksi, dr = 0.33, 0.7-in. diameter stands, N ≈ 0.25Nmax)
c.  U-girders: 10 No. 6 Gr. 60 Texas U40
(L = 95 ft, f c′ = 12 ksi, dr = 0.68, 0.7-in. diameter strands, N ≈ 0.50Nmax)
7. The box girders required the fewest amount of additional nonprestressed reinforcement;
25 cases, which correspond to 11% of the total number of box girders, required nonprestressed
reinforcement. In the case of I girders and U-girders, nearly the same percentage of the total
cases required some additional nonprestressed reinforcement: 32% (55 cases) for I girders
and 33% (44 cases) for U-girders. While No. 4 Gr. 60 reinforcing bars sufficiently augmented
the tensile capacity of I girders with partially debonded strands, No. 6 Gr. 60 reinforcement
had to be provided for a number of U-girders.

2.4  Finite Element Method Modeling


An extensive FEM modeling program was developed in this study with the goal of developing
an analytical platform from which a deeper understanding of the local and global behaviors of
prestressed girders containing partially debonded strands could be established. The resulting
3D FEM platform significantly extended the parametric scope of the experimental study, which
was limited by time, financial resources, and specimen size.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

22   Strand Debonding for Pretensioned Girders

Commercial software was used to conduct all 3D FEM analyses. The software is focused on
reinforced concrete structures and is widely used for both design and research. A summary of
the FEM platform features is described in the following sections. A detailed description of the
modeling technique and validation studies conducted is provided in Appendix C.

2.4.1  Development and Validation of 3D-FEM Model


2.4.1.1  Material Models
A fracture-plastic model was employed to describe the behavior of concrete. The model—
consisting of a fracture model based on a smeared crack formulation and crack band concept,
and a plasticity model residing on a plastic failure surface—is intended to accurately capture
concrete behavior in tension (fracture) and compression (crushing). By combining fracture and
plasticity models, tensile strength reduction after concrete crushing and compressive strength
reduction after concrete cracking are both taken into account. Additionally, the shear strength
of cracked concrete was calculated based on the modified compression field theory (Vecchio
and Collins 1986). Tension stiffening, based on the CEB-FIP model (1990), was also considered
in the constitutive model used. Concrete modeling parameters were based on default values or
values obtained from AASHTO-prescribed formulas. The only free parameter in the modeling
conducted was the compressive strength of concrete.
For nonprestressed reinforcement, the classical elasto-plastic material model was employed.
Additionally, a rigid connection was assumed between concrete and reinforcement (i.e., “perfect
bond” condition).
For prestressing strands, a more accurate representation of the strand response, based on a
Ramberg-Osgood model, was used to describe the nonlinear stress-strain relation. For strands,
perfect bond is not realistic; indeed, bond slip can initiate immediately upon prestress transfer.
A bond stress-slip model following the formulation of the CEB-FIP model (1990) was used to
describe the mechanical interaction at the concrete-strand interface. The bond model was scaled
such that it would result in prescribed transfer lengths of either a “realistic” value of 30db or the
AASHTO-prescribed value of 60db (shown in Figure 2.5). The bond model calibrations were

Figure 2.5.   Representative example of pattern of bonded strands


and longitudinal distribution in AASHTO Type IV girder—Case 1 shown.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Analytical Research Approach and Findings   23  

validated and a mesh-sensitivity study conducted using experimental data from Burgueño and
Sun (2011), as described in Appendix C.

2.4.1.2  Structural Modeling


In each model, the concrete of the girder was meshed with 3D hexahedral elements, while the
reinforcement and prestressing strands were modeled with 3D truss elements. Their interaction
is captured by the bond-slip model based on the relative displacement between the steel truss
element and the surrounding hexahedral concrete element.
The prestress of the strands was modeled in a manner similar to the practice of pre-tensioning.
An external force required to generate the desired prestress was applied to each strand. After
concrete was “cast” (concrete modeled with properties corresponding to those at prestress
transfer), the external force was deactivated to simulate the prestress transfer. The construction
of the slab was modeled similarly. Initially, the top slab was “deactivated” during girder construc-
tion so that it would have no effect at prestress transfer. Following this step and an increase in
the underlying girder concrete strength, the top slab load was first applied and then the slab
was “activated” to work together with the girder to resist the simulated AASHTO-prescribed
STRENGTH I and SERVICE I and III loadings. In this manner, the critical loading stages for the
prestressed girder were modeled in a realistic manner.
In addition to output files, the commercial software provides a post-processing interface to
graphically illustrate the simulation results. The stress and strain in concrete and strands can
be graphically demonstrated on the model, as well as the cracks shown. The software permits a
threshold level of damage to be illustrated; that is, only damage (crack widths) greater than this
threshold is shown in output graphics. This feature of the analysis software was used extensively
in this report.
Two FEM models of previous laboratory-tested girders (Hawkins and Kuchma 2007) were
used to validate the FEM model. All FEM material properties and geometries were consistent
with those reported, and the actual transfer length of 38db, determined from elastic shortening
measurements, was used. In order to accurately capture the girder construction process, the
analysis was divided into four loading phases: (1) casting girder (analysis step 1), (2) prestressing
strand release (steps 2–7), (3) cast top slabs (step 8), and (4) external loading as described in
Hawkins and Kuchma (steps 9–26). Details of the experimental girders and FEM modeling are
provided in Appendix C.
As shown in Appendix C, ultimate capacity and failure mode, load-deflection behavior at
various locations along the spans, and crack patterns all indicated excellent agreement between
the FEM-predicted and experimental behavior. In particular, the FEM model was able to capture
the splitting failure associated with partial debonding of one girder.

2.4.2  FEM Parametric Study


A parametric FEM study, aimed at accumulating the following information important to
structural analysis and safety evaluation of prestressed girders with partially debonded strands,
was conducted:
1. Crack patterns and stress distribution at the critical loading stages of prestress transfer, service
conditions, and at ultimate load;
2. Effects of partial debonding on crack patterns and failure modes; and
3. The impact of local effects such as the debonding details and bond interface properties.
A matrix of 44 3D models, shown in Table 2.3, was selected based on the outcome of the larger
parametric design study described in Section 2.3 and Appendix B. Five girder shapes: AASHTO

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

24   Strand Debonding for Pretensioned Girders

Table 2.3.   Cases for 3D FEM analytical study.


f’c AASHTO Type IV Nebraska NU-900
(ksi) Case Strands dr L (ft) Case Strands dr L (ft)
8 14-0.6 0.14 55
6 - - -
9 14-0.7 0.43 65
10 30-0.5 0.33 70
8 1 34-0.5 0.59 85 11 14-0.6 0.00 55
12 14-0.7 0.29 65
2 66-0.5 0.73 115 13 60-0.5 0.53 100
12 3 50-0.5 0.64 100 14 44-0.5 0.36 85
4 34-0.5 0.47 85 15 30-0.5 0.07 70
16 60-0.5 0.47 100
5 66-0.5 0.70 115
17 44-0.5 0.27 85
15 6 50-0.5 0.60 105
18 30-0.5 0.00 70
7 34-0.5 0.41 85
19 60-0.6 0.67 115
AASHTO BIV-48 adjacent Texas U-54
20 46-0.5 0.57 120
6 21 34-0.5 0.41 105 33 42-0.5 0.45 80
22 23-0.5 0.13 85
23 46-0.5 0.52 120
34 63-0.5 0.57 95
8 24 34-0.5 0.36 105
35 42-0.5 0.36 80
25 46-0.6 0.65 140
26 46-0.5 0.35 120
36 85-0.5 0.60 110
27 34-0.5 0.12 105
12 37 63-0.5 0.46 95
28 46-0.6 0.57 145
38 42-0.5 0.19 85
29 46-0.7 0.70 165
39 85-0.5 0.53 115
30 46-0.5 0.22 120
40 63-0.5 0.37 95
15 31 46-0.6 0.52 145
41 42-0.5 0.05 85
32 46-0.7 0.65 165
42 85-0.6 0.71 135
AASHTO BT-72
43 38-0.7 0.63 135
15
44 28-0.7 0.50 115
dr = debonding ratio

Type IV, Nebraska NU-900, AASHTO BIV-48 (in adjacent box arrangement), Texas U-54,
and BT-72 girders were considered. The effects of span, extent of partial debonding, concrete
strength, and strand size on the structural performance of the girders were explored. For each
girder, three critical loading phases were studied: prestress transfer, service (AASHTO SERVICE
I and III), and ultimate (STRENGTH I) limit states. SERVICE and STRENGTH limit states
considered live load arrangement intended to maximize either flexure or shear. Additionally,
each model was “loaded to failure” (in both flexure and shear) to assess its ultimate capac-
ity. A summary of key stress checks made in this study is presented in Table 2.4. A summary
of the results of each analysis is presented in single-page summary matrices provided in
Appendix D.

2.4.2.1  Modeling Parameters


For consistency, a number of parameters and details are kept constant or consistent across all
analyses. These parameters are described as follows:

2.4.2.1.1  Concrete Strength.    Concrete strength, f c′, is given in Table 2.3 for each model. In
all cases, the concrete strength at prestress transfer was taken as fci = 0.6 f c′, a conservative lower
′ = 5 ksi, a value representative of in situ
bound. The slab strength in every case is taken as f c,slab
slab strength.

2.4.2.1.2  Prestressing Strand.    Strand diameter, db, is given in Table 2.3 for each model. In
all cases, 270-ksi low-relaxation strand was used. Initial prestress was taken as 0.75fpu = 202.5 ksi
in every case. Long-term prestress accounting for all losses (but not bond slip, which is accounted

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Analytical Research Approach and Findings   25  

Table 2.4.   Stress criteria checked in 3D FEM models.


Location Criteria At Prestress STRENGTH I SERVICE I and III
Along Transfer
Girder
Length
Concrete
ft 0.24 f’ci 1 Not checked Not checked
Near tension
support T Apsfps
(At support HL-93 loading for shear
Prestressing OK if satisfied for
face, dv/2 Not checked strand is not developed at
steel STRENGTH I
and dv) this location; thus, fps <
fpu
At SERVICE III
Concrete
ft 0.24 f’ci 1 Not checked ft 0.19 f’c
tension
HL-93 loading for flexure
At SERVICE I
Concrete fc 0.45 f’c permanent load
At midspan fc 0.60f’ci Not checked
compression fc 0.60 f’c
HL-93 loading for flexure
OK if satisfied T Apsfpu OK if satisfied for
Prestressing
for STRENGTH HL-93 loading for flexure STRENGTH I
steel
I
1
It will be additionally noted if ft ≤ 0.0948 f’ci ≤ 0.2 ksi.

for directly in the model) was taken as 0.56fpu = 151.2 ksi in all cases. Strand bond parameters
were calibrated such that the transfer length would be 60db.

2.4.2.1.3  Partial Debonding.    The maximum partial dr for each model is given in Table 2.3.
This value is the dr at the girder end. Partially debonded strands were introduced (bonded)
into the cross section in three approximately equal increments of 3 ft each. Thus, at a location
9 ft into the span, all strands in the section were bonded. Figure 2.5 shows an example for Case 1,
an 85-ft long Type IV girder having 34 strands, 20 of which are partially debonded (dr = 0.59).
The debonding pattern used in each analysis is summarized in the respective summary matrix
(Appendix D) and is generally consistent (except in cases with very large debonding ratios) with
the proposed detailing guidelines (Chapter 4). The overall girder behavior was not significantly
affected by the strand debonding pattern used; strand debonding patterns primarily affected
local transverse stresses in the bulb as discussed in Section 2.5.

2.4.2.1.4  Shear Reinforcement.    Shear reinforcement was modeled as discrete bars in all
cases and included both vertical web reinforcement (extending into the slab) and bulb or flange
confinement reinforcing appropriate for the girder shape. All shear reinforcement was assumed
to have a yield strength of fy = 60 ksi. Shear reinforcement details are given in each summary
matrix (Appendix D).

2.4.2.1.5  Boundary Conditions.    All girders were modeled as simply supported beams
having a full-width 12-in. long rigid bearing supported by a pin at one end and a roller at the
other. Full width is understood to mean the full width of the bottom flange less a distance
accounting for the chamfer, typically 2 in. on both sides. This arrangement allowed realistic rota-
tion at the girder end during prestress transfer and reasonably mimics neoprene-bearing pads,
once the girder would be placed in service.
The entire cross section of all girders, except BIV-48, was modeled. A half section of BIV-48
was modeled having boundary conditions along the line of symmetry that enforce the assumed
plane sections behavior of the box section.

2.4.2.1.6  Applied Loads.    In all the analyses, permanent loads (DC) included girder self-
weight, slab weight, and an additional 300 lb/ft to account for barrier wall loads (two 750 lb/ft

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

26   Strand Debonding for Pretensioned Girders

walls distributed over five girders). An additional wearing surface load (DW) is taken as 35 psf
(3-in. thick asphalt overlay). HL-93 live loading (LL), considering appropriate distribution
factors and vehicle placement for moment or shear, was used in all analyses. AASHTO impact
loads (IM) were included in all the appropriate cases.

2.4.2.1.7  Modeling Steps.    As described previously, modeling followed the typical order of
prestressed concrete construction. Six primary steps are summarized in Table 2.5. Release of
tendons and application of loads were introduced over a number of substeps to permit load
redistribution. No inertial properties were modeled; therefore, tendon release was assumed to
be “slow,” having no impact or dynamic effects.
Application of vehicle loads in Steps 5 and 6 were repeated with the vehicle located on the span
to maximize the effects for either flexure or shear, consistent with AASHTO design requirements.
In both cases, Step 6 incrementally increased the HL-93 axle loading above the STRENGTH I
limit state condition (Step 5) until girder failure. Therefore, the value of a shown in Table 2.5
was greater than 1.75 (the STRENGTH I limit state). Due to the complexity of modeling the loads,
the lane portion of the HL-93 was not increased above the STRENGTH I limit state (i.e., the lane
load was “constant” having an applied load factor of 1.75).

2.4.2.1.8  Model Conventions.    All FEM models simulated the entire girder span, the
appropriate effective slab width, and were symmetric about midspan. The origin was defined
at the midspan soffit. Thus, in all figures showing only half the span, the axis labels indicate the
midspan of the girder. All vehicle loads were applied such that the left end (as shown in figures)
of each girder is critical. When half spans are shown, it is the west end of the girder and the
longitudinal profiles are not to scale: the vertical dimension was stretched to enhance clarity.
Figure 2.6 illustrates the details of representative commercial software models for each girder
shape considered.

2.4.2.2  Simulation Summary


Detailed model information, stress distributions, crack patterns, and local responses of each
girder model are provided as single-page summary matrices in Appendix D. A summary of all
checks (described in Table 2.4) is provided in Table 2.6. Values shown in bold text indicate that
the value fails the check; otherwise, all stresses and ratios presented pass the AASHTO-prescribed
stress checks. The stress checks summarized in Table 2.6 are as follows.

2.4.2.2.1  Prestress Transfer.    The tension (T) check verifies that at no location along the
girder does the concrete tensile stress exceed the AASHTO-prescribed limit, that is, ft ≤ 0.24√ f ci′
[in this case, ft ≤ 0.24√(0.6 f c′)]. This limit is usually associated with cracking at the top surface of the

Table 2.5.   FEM modeling steps.


Step Description Applied Loads Concrete External Internal
Strength Prestress Prestress
1 “Cast” concrete None N.A. 0.75fpu N.A.
2 Release tendons Girder only fci = 0.6f’c 0 0.9(0.75fpu)1
3 Place slab Girder and slab f’c 0 0.56fpu
4 SERVICE I DC+DW+(LL+IM) f’c 0 0.56fpu
SERVICE III DC+DW+0.8(LL+IM)
5 STRENGTH I 1.25DC+1.50DW+1.75(LL+IM) f’c 0 0.56fpu
Flexure and shear
6 Failure 1.25DC+1.50DW+ 1.75(LLlane) f’c 0 0.56fpu
+ (LLtruck+IM)
Flexure and shear
1
Losses upon transfer were determined within FEM model based on the provided bond slip
model; in general, losses were approximately 10% of the initial prestress force.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Analytical Research Approach and Findings   27  

(a) Type IV girder

(b) NU-900 girder

(c) BIV-48 girder (half-section model)

(d) U-54 girder

(e) BT-72 girder

Figure 2.6.   Representative ATENA models showing complete girder (left) and west half span and
section (right).

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Table 2.6.   Summary of FEM simulation stress checks and predicted ultimate capacity.
Stress Checks (See Table 2.4)
Ultimate
At Prestress
Model Parameters (See Table 2.3) SERVICE STRENGTH I Capacity
Transfer
T C I III Shear Flexure Shear
Flexure
f’c db N L dr S ft fc ft fc Aps fps /T at section:
Case Cracking Cracking
ksi in. no. ft - ft ksi ksi ksi ksi Support dv/2 dv See App. D
AASHTO Type IV girders
1 8 0.5 34 85 0.59 0.28 2.16 0.21 1.50 minor none 0.79 1.01 1.12 3.15 4.75
2 12 0.5 66 115 0.73 0.39 4.00 0.35 2.83 none none 0.74 1.02 1.09 3.75 4.75
3 12 0.5 50 100 0.64 0.38 3.38 0.33 2.30 none none 0.70 1.01 1.11 3.05 5.55
4 12 0.5 34 85 0.47 8 0.43 2.45 0.26 1.53 minor none 0.74 1.04 1.15 2.95 5.35
5 15 0.5 66 115 0.70 0.42 4.17 0.36 2.87 none none 0.71 1.02 1.10 3.45 5.15
6 15 0.5 50 105 0.60 0.47 3.34 0.31 2.27 minor none 0.70 1.03 1.11 3.05 4.75
7 15 0.5 34 85 0.41 0.53 2.51 0.30 1.62 minor none 0.73 1.05 1.14 3.00 5.35
Nebraska DOT NU-900 girders
8 6 0.6 14 55 0.14 0.33 1.56 0.17 1.12 some minor 0.63 0.89 0.99 2.75 3.15
Case 8 with additional As =
8A 0.31 1.58 0.17 1.12 some minor 0.69 0.98 1.08 2.75 3.20
0.88 in.2
9 6 0.7 14 65 0.43 0.23 1.94 0.15 1.01 some minor 0.62 0.82 0.96 1.95 3.75
Case 9 with additional As =
9A 0.23 1.99 0.15 1.01 some minor 0.71 0.99 1.10 2.00 3.80
1.32 in.2
10 8 0.5 30 70 0.33 0.26 2.16 0.28 1.31 some none 0.77 1.02 1.04 2.55 3.35
11 8 0.6 14 55 0.00 0.28 1.85 0.21 1.31 some minor 0.65 0.93 1.01 2.15 3.35
8
Case 11 with additional As =
11A 0.30 1.78 0.21 1.31 some minor 0.69 0.99 1.07 2.15 3.45
0.62 in.2
12 8 0.7 14 65 0.29 0.32 2.03 0.18 1.26 some minor 0.64 0.81 0.96 2.55 3.75
Case 12 with additional As =
12A 0.31 2.08 0.18 1.26 some minor 0.74 1.01 1.10 2.55 3.75
1.32 in.2
13 12 0.5 60 100 0.53 0.31 3.73 0.58 2.40 some some 0.74 1.00 1.02 2.55 3.35
Case 13 with all debonded
13B strands introduced at one 0.26 3.90 SERVICE and STRENGTH load cases not run
location (6 ft)
14 12 0.5 44 85 0.36 0.33 3.01 0.43 1.90 some none 0.70 1.01 1.05 2.15 3.55
15 12 0.5 30 70 0.07 0.35 2.34 0.31 1.82 some none 0.69 1.02 1.06 2.15 3.55
16 15 0.5 60 100 0.47 0.36 3.81 0.72 2.45 some none 0.70 1.02 1.03 2.15 4.15
17 15 0.5 44 85 0.27 0.39 3.06 0.45 2.08 some none 0.69 1.02 1.04 2.15 3.75
18 15 0.5 30 70 0.00 0.38 2.86 0.33 2.00 some none 0.62 1.03 1.07 2.15 3.15
19 15 0.6 60 115 0.67 0.36 5.07 0.51 4.80 minor minor 0.75 1.03 1.09 2.35 3.85
AASHTO BIV-48 adjacent box girders
20 6 0.5 46 120 0.57 0.16 2.67 0.08 2.22 none none 0.72 1.02 1.05 3.55 6.15
21 6 0.5 34 105 0.41 0.17 2.03 0.09 1.77 none none 0.72 1.01 1.09 3.95 6.55
22 6 0.5 23 85 0.13 0.19 1.82 0.14 1.52 none none 0.70 1.00 1.12 3.15 4.35
23 8 0.5 46 120 0.52 0.26 2.80 0.10 2.23 none none 0.72 1.01 1.06 4.75 5.75
24 8 0.5 34 105 0.36 0.23 2.10 0.12 1.70 none none 0.72 1.02 1.10 3.75 7.35
25 8 0.6 46 140 0.65 0.22 3.70 0.14 2.30 none none 0.79 0.99 1.01 4.95 7.15
26 12 0.5 46 120 0.35 0.33 2.89 0.15 2.26 none none 0.71 1.02 1.09 3.75 8.75
27 12 0.5 34 105 0.12 0.32 2.27 0.18 1.89 none none 0.70 1.02 1.14 3.15 6.15
28 12 0.6 46 145 0.57 4 0.30 3.93 0.13 3.24 none none 0.65 0.98 1.00 4.55 8.75
29 12 0.7 46 165 0.70 0.30 5.24 0.21 3.48 none none 0.68 0.90 0.95 5.15 7.35
Case 29 with additional As =
29A 0.22 5.24 0.21 3.48 none none 0.75 1.01 1.05 5.20 7.35
1.32 in.2
30 15 0.5 46 120 0.22 0.38 2.91 0.20 2.52 none none 0.71 1.03 1.13 3.55 7.15
31 15 0.6 46 145 0.52 0.42 4.04 0.18 3.26 none none 0.76 0.98 1.04 4.15 6.55
Case 31 with additional As =
31A 0.30 4.20 0.18 3.26 none none 0.78 1.00 1.07 4.15 6.55
0.44 in.2
32 15 0.7 46 165 0.65 0.35 5.25 0.19 3.54 none none 0.64 0.89 0.97 5.55 8.35
32A Case 32 with additional As = 0.26 5.48 0.19 3.54 none none 0.70 1.00 1.07 5.50 8.35
1.32 in.2
Texas DOT U-54 girders
33 6 0.5 42 80 0.45 0.26 1.85 0.14 2.36 some minor 0.69 1.00 1.12 2.55 4.75
34 8 0.5 63 95 0.57 0.29 2.67 0.20 3.19 some none 0.67 1.01 1.14 3.35 5.00
35 8 0.5 42 80 0.38 0.29 1.89 0.17 2.31 some minor 0.71 1.00 1.13 2.75 4.75
36 12 0.5 85 110 0.60 0.32 3.68 0.19 4.00 minor none 0.73 1.01 1.14 3.35 5.60
37 12 0.5 63 95 0.46 0.35 2.69 0.31 3.16 some none 0.76 1.01 1.13 2.35 4.75
14
38 12 0.5 42 85 0.19 0.36 2.01 0.51 2.48 some some 0.73 1.01 1.16 1.75 3.35
39 15 0.5 85 115 0.53 0.46 3.73 0.21 4.27 some none 0.75 1.01 1.15 2.95 5.35
40 15 0.5 63 95 0.37 0.41 2.71 0.23 3.18 minor none 0.72 1.01 1.13 3.15 5.55
41 15 0.5 42 85 0.05 0.41 2.30 0.62 2.89 some some 0.70 1.01 1.18 1.95 2.75
42 15 0.6 85 135 0.71 0.42 4.98 0.23 6.09 minor none 0.69 1.00 1.00 2.75 5.55
BT-72 girders
43 15 0.7 38 135 0.63 0.32 5.38 0.26 4.43 minor minor 0.69 1.05 1.14 5.45 7.50
8
44 15 0.7 28 115 0.50 0.35 4.92 0.29 3.18 some minor 0.67 1.03 1.12 5.50 8.80

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Analytical Research Approach and Findings   29  

girder near the girder end, as is clearly shown in the crack patterns at prestress transfer shown
in Appendix D. For all cases considered (except 20 and 21) 0.0948√ f ci′ < ft ≤ 0.24√ f ci′ ; therefore,
nonprestressed reinforcement satisfying the requirements of AASHTO LRFD Table 5.9.4.1.2-1
is required in the region of tensile stress. All cases shown in Table 2.6 satisfy the ft ≤ 0.24√f ci′
limit; indeed the drs were selected to ensure this requirement would be met. It is noted that the
required additional tensile steel needed to control cracking was not provided in the models
since the intent of this study was to assess the effectiveness of the debonding alone to mitigate
these tension stresses.
The compression (C) check verifies that at no location along the girder does the concrete com-
pressive stress exceed the AASHTO-prescribed limit, that is, fc ≤ 0.60f ci′ (in this case, fc ≤ 0.36 f c′).
As indicated in Table 2.6 by bold entries, a few BIV girders fail this check; these cases would
likely be considered impractical spans, but were included in this study in order to capture the
full range of potential behavior. In each case that failed this check, considering a higher concrete
strength at prestress transfer would result in the stress limit being satisfied. For example, Case 23
( fc = 2.80 ksi ≤ 0.60 f ci′ ; where f ci′ = 3.6 ksi) closely replicates Case 20 ( fc = 2.67 ksi > 0.60 f ci′ ;
where f ci′ = 4.8 ksi) but with a higher concrete strength; the compressive stress check passes in
the latter case.
Images of the cracking at prestress transfer are shown in the individual simulation summaries in
Appendix D. A crack threshold of 0.000 in. is used to illustrate cracking; therefore, any predicted
cracking, regardless of size, is indicated. The maximum crack size observed is also noted.

2.4.2.2.2  SERVICE I Limit State.    The SERVICE I check verifies that at no location along the
girder do the concrete tensile stresses exceed the AASHTO-prescribed limit, that is, ft ≤ 0.19√ f c′.
The SERVICE III check verifies that at no location along the girder do the concrete compressive
stresses exceed the AASHTO-prescribed limits, that is, fc ≤ 0.45f c′ under the effects of permanent
load and fc ≤ 0.60 f c′ under the effects of HL-93 loading for flexure. All girders performed adequately
under SERVICE I and SERVICE III loading.

2.4.2.2.3  STRENGTH I Limit State.    The primary consideration at the STRENGTH I limit
state is the performance of the strands. Since the girder spans were selected based on satisfying
moment capacity of the section, all sections perform adequately at the section under maximum
moment. Similarly, since the spans were intentionally stretched in this study to maximize the
utilization of a section, shear capacity is also adequate. The focus of this study was to evaluate
the ability of the strands to develop the required tensile forces near the girder supports. The
criterion, in this case, is that the tensile force, T, developed in the strands exceeds that available:
Aps fps. Near the supports, debonded strands are not included in the Aps term and strands that have
not been fully developed may only contribute a strand force fps that is less than fpu. In Table 2.6,
this check has been expressed as the calculated ratio Aps fps /T. A value exceeding unity indicates
that the check is satisfied. This check is made at three sections: the face of the support, and at
distances of dv /2 and dv from the face of the support, where dv is the shear depth of the member.
The distance dv from the face of the support is considered the critical section for shear (AASHTO
LRFD Article 5.8.3.5).
As indicated in Table 2.6 by bold entries, all girders fail this check at the face of the support
although most satisfy this requirement at a distance of dv /2 from the support face. In the FEM
analyses conducted, which captured the effects of strand slip, this behavior indicates that while
there may be localized strand slip in the strand transfer length, adequate residual bond remains
to develop the strand further along its length. That is, the effect of the strand bond capacity
being exceeded at the face of the support does not lead to failure of the girder; indeed most
girders demonstrated significant reserve shear capacity beyond the STRENGTH I limit state as
described below.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

30   Strand Debonding for Pretensioned Girders

Some BIV and NU-900 sections also fail to meet the Aps fps /T ≥ 1.0 criterion at dv /2 and dv.
Despite not meeting the criterion, redistribution is adequate such that significant reserve capacity
is still available. Most girders meet the strand tension requirement at dv. Since the girder details
were selected to maximize capacity while still generally meeting strand tension limits, the results
are a vindication of the design approach used to maximize the girder spans in the first place.
For those BIV girders in which Aps fps /T < 1.0, the debonding ratios are all quite high: dr = 0.52,
0.57, 0.65 (two cases), or 0.70. For the NU girders that fail this check, however, the debonding
ratios are generally lower, some respecting current AASHTO limits (dr < 0.25): dr = 0.00, 0.14,
0.29, 0.43, or 0.67.
In the FEM study, the support lengths were taken as only 12 in. and no overhang behind
the support (allowing a longer transfer length) was provided. Methods of mitigating the low
tensile capacity at the support face include providing supplemental nonprestressed reinforcing
steel (As fy in AASHTO LRFD Eq. 5.8.3.5-2) or providing additional anchorage for the strand
in either an overhang or by embedding it in an integral abutment or diaphragm (effectively
increasing fps).

2.4.2.2.4  Ultimate Capacity.    Following the STRENGTH I limit state, the HL-93 axle loads
were increased. The value of a shown in Table 2.6 corresponds to the live load factor to cause
girder failure, i.e., 1.25DC + 1.50DW + 1.75(LLlane) + a(LLtruck + IM). The values for critical
flexure and shear are provided for all cases. Since girder span length was stretched, it should be
expected that all girders are “flexure critical,” that is, the value of a causing failure is lower for
flexure than for shear.
The predicted ultimate behavior for all girders subjected to flexural-critical loading is an
expected flexural mode of failure. Most analyses indicate prestressing strand strains in excess
of 1%. Crushing of the slab is evident primarily in girders having lower concrete strength (the
lower girder modulus leads to greater top flange compression strains and, therefore, crushing
of the lower-strength composite slab). Ten of the 44 girders demonstrated flexure-shear failures
in which, typically near girder ends, shear distress accompanies the dominantly flexural tension
failure. Girders that exhibited flexure-shear failures generally had higher debonding ratios
(0.14, 0.29, 0.33, 0.43, 0.53 [twice], 0.60, 0.67, 0.70, or 0.71); thus, the strand tension capacity, T,
likely played a role in the shear component of the predicted flexure-shear failure. Considering
the NU girders, all of those that failed the Aps fps /T ≥ 1.0 check at the STRENGTH I limit state also
demonstrated flexure-shear failures. None of the BIV girders, however, demonstrated significant
shear distress in their ultimate behavior, perhaps due to the proportionally larger web area of
such sections (especially compared to the thin-webbed NU section).
The ultimate behavior under shear loading is more informative and appears to identify a
few trends. As shown in Figure 2.7, four primary modes of failure were observed (top down in
Figure 2.7b and c): (1) single dominant shear band or strut associated with an axle load (19 cases),
(2) distributed shear cracking (10 cases), (3) flexure-shear behavior (6 cases), and (4) flexural
tension or strand yield (9 cases).
As seen in Figure 2.7b, shear failures are dominant in girders having greater debonding ratios.
Additionally, higher debonding ratios tended toward the formation of single crack bands. Shear
failure corresponds to a failure of the Aps fps /T ≥ 1.0 criterion at the critical section for shear.
Nonetheless, one must be careful with this conclusion because girders having higher debonding
ratios are naturally those with a higher prestress reinforcing ratio (r = NAps /Ac) (Figure 2.7a). Such
girders are “over-reinforced” for flexure and, therefore, are expected to have a shear-dominated
behavior. The corollary of this observation is that those girders having a low prestress reinforcing
ratio (and therefore lower debonding requirements) tend to fail in a flexural mode of behavior.
These trends are exhibited graphically between Figures 2.7a and b.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Analytical Research Approach and Findings   31  

0.8
Case 42
0.7
Single shear band
0.6
Case 13
Debonding ratio, dr

0.5
Distributed shear
0.4
Case 37
0.3
Flexure-shear
0.2
Case 38
0.1
Flexural tension
0.0
0.00 0.01 0.02 0.0 0.2 0.4 0.6 0.8
Prestress reinforcing ratio, Debonding ratio, dr
ρ = Aps /Ag
(a) Greater prestress reinforcing (b) Debonding ratio and failure mode (c) Representative failure modes
ratio results in greater (horizontal and vertical scales vary)
debonding ratio

Figure 2.7.   Qualitative relationship between prestress reinforcing ratio, dr, and ultimate failure modes
for cases having no additional nonprestressed reinforcement (i.e., no cases labeled A in Table 2.6).

2.4.2.2.5  Performance of Different Girder Cross Sections.    Little difference was observed
in the overall behavior of AASHTO Type IV and NU-900 girders. Generally, NU girders
when “stretched” to their longest practical spans exhibited lower concrete stresses at prestress
transfer—indicating less need for debonding. NU girders also had proportionally lower ulti-
mate capacities and therefore exhibited greater cracking at the STRENGTH I limit state. These
comparisons reflect the “more efficient” section of the NU family of girders as compared to the
AASHTO type girders. The NU-900 girders that failed the Apsfps/T ≥ 1.0 criteria had few larger
diameter strands near their support. The larger strands have longer transfer lengths and, hence,
the fps term is developed more gradually along the girder span. In each of the cases, the Aps term is
also relatively low near the support. Both Type IV and NU-900 girders exhibited limited evidence
of longitudinal web cracking at the girder ends resulting from prestress transfer. Such cracking is
occasionally observed in the field and is associated with large prestress forces (Kannel et al. 1997).
The BIV-48 girders behaved well—illustrating no cracking exceeding 0.016 in. at the
STRENGTH I limit state and proportionally larger ultimate capacities. Due to the greater com-
bined web dimension, ultimate shear capacity was quite high. Stresses at prestress transfer for
the BIV-48 girders were consistent with all other girders considered.
The BIV-48 girders that failed the Aps fps /T ≥ 1.0 criteria did so for a very different reason than
the NU-900 girders. The BIV girders that failed were all very long (between 140 and 165 ft)
resulting in relatively large flexural demands. Due to their length, the shear load case results in
minimal flexural demand; the shear capacity is, therefore, quite high, driving up the strand tension
demand, T. The BIV-48 girders that failed the Aps fps /T ≥ 1.0 criteria all had very high ultimate shear
capacities in which the value of a exceeded 7.

2.5  Strut-and-Tie Modeling of End Regions


It is important to note that although the issue of longitudinal splitting described was identified
and initially associated with a case of strand debonding (Section 2.5.1), the issue is not restricted to
girders having debonded strands. Rather, the contents of this section must be interpreted to apply
to the pattern of bonded strands at a girder end, irrespective of the presence of debonded strands.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

32   Strand Debonding for Pretensioned Girders

2.5.1  Motivating Example


Partial debonding of strands is required to reduce concrete stresses and mitigate cracking in
girder end regions. Nonetheless, the details associated with partial debonding can, in fact, cause local
cracking and potentially lead to early and abrupt ultimate-limit-state failure, particularly at girder
ends (Ross et al. 2013). This effect was illustrated in one test specimen reported by Hamilton (2009)
in a study focusing on enhancing the shear performance of existing girders using fiber-reinforced
polymer (FRP) materials. Although the Hamilton study is beyond the scope of the reported research,
a critical serviceability issue was identified. Hamilton (2009) considered specimens designed to
replicate in-service Type IV girders. Of interest is the fact that the partial debonding provided
violated Article 5.11.4.3 of the AASHTO LRFD Bridge Design Specifications (2012) by having 57%
of all strands debonded, as seen in Figure 2.8b. Additionally, the debonded strands were all in
the middle of the section and the 18-in.-wide bearing pad did not extend beneath the remaining
bonded strands (Figure 2.8). As a result, the compressive strut resulting from the application of
a shear load was required to spread from the web to engage the outermost bonded strands and
“return” back inward to be reacted at the bearing pad. This transfer necessitates the formation of a
transverse tie as shown schematically in Figure 2.8c. As can be seen in Figure 2.8a, the presence of
this tie resulted in vertical cracks through the bulb. Additionally, the compressive thrust appears
to have “cracked out” or “pushed off” the cover concrete on the left side of the bulb (Figure 2.8a).

2.5.2  STM of Prestressed Girder End Region


Ross et al. (2013) proposed an STM approach, as an alternative to extant prescriptive detail
requirements, for designing bulb confinement reinforcement to mitigate lateral splitting failures
at the ultimate limit state. The STM proposed by Ross et al.(2013) is similar in geometry to that
proposed here but with significantly greater simplification that makes the investigation of the
effects of strand pattern not feasible. In the present work, no such simplification is made; in fact,
the analyses conducted for this project uses what Ross et al. (2013) ultimately recommended:
“. . . the strut angle must be calculated directly.”
An STM provides a simple yet effective tool to investigate the local effects of the pattern of
bonded strands (and, by extension, debonding pattern) on cracking at the ends of girders. The
following assumptions are typical of the STM approach and permit the 3D STM at the end of
the girder (Figure 2.8c) to be modeled as a two-dimensional STM as shown:
1. All sections and STMs are symmetric about the vertical centerline of the girder cross section.
2. The horizontal strands are sufficient to anchor the inclined strut in the longitudinal direction of
the girder. With this assumption, the angle of this strut does not affect the 2D (cross sectional)
STM from capturing the section behavior.

(a) Details and cracking (b) Partially debonded strands (c) STM

Figure 2.8.   Lower bulb of Type IV girder (photos: Hamilton 2009).

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Analytical Research Approach and Findings   33  

3. Struts and ties are anchored at nodes corresponding to centroids of groups of bonded pre-
stressing strand. This assumption implies that all bonded strands resist the generated tie force
equally, which is consistent with the Bernoulli beam assumption (i.e., plane sections remain
plane). Extending this assumption to the cross section STM, each strut will be anchored at
a node corresponding to the centroid of a strand group and the corresponding strut force is
proportional to the number of bonded strands represented by the node.
4. Girders are not skewed at their ends.
The proposed STM adopted in the present study is shown in Figure 2.9.
In Figure 2.9, the following nomenclature is used:
B2, B3, D5, and D6 are girder dimensions (PCI 2011).
Vu = total reaction (shear) at support.
Nw = total number of bonded strands at section.
nf = number of bonded strands in one side of the outer portion of bulb. The outer portion
of bulb is defined as that extending beyond projection of web width, B3. Strands aligned
with the edge of web are assumed to fall in the outer portion of bulb.
xp = horizontal distance to girder centerline of centroid of nf strands in outer portion of bulb.
yp = vertical distance to girder soffit of centroid of nf strands in outer portion of bulb.
cb is calculated to ensure uniform bearing pressure across width of bearing pad, bb, i.e., Eq. 2.8:

cb = (bb 2 )(1 − n f Nw ) Eq. 2.8

Although a uniform distribution is used here, the calculation of cb may be refined to reflect any
distribution of bearing stress across the bearing pad such as would be the case where the bearing
was treated as a Winkler beam.
The tension in the horizontal tie, t, located at yp from the soffit may be calculated (Eq. 2.9):

t = αVu φ = (n f N w )[ x p ( hb − yp ) + ( x p − cb ) yp ]Vu φ Eq. 2.9

The tie force is written as a fraction, a, of the reaction force. This equation is essentially iden-
tical to the Ross et al. (2013) equation but with cb, xp, and yp calculated based on strand pattern

Figure 2.9.   STM (shown on BT girder bulb).

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

34   Strand Debonding for Pretensioned Girders

and girder geometry, whereas Ross et al. (2013) made the following simplifications: yp = D6/2 and
xp = cb= (B2 - B3)/3. An additional tension tie resisted at the level of the bearing pad is required
when xp - cb is negative. Such cases only arise for small nf /Nw and, thus, are unlikely to result in
large tension tie forces. It is assumed that the bearing itself will resist this tie.
Based on sloped geometry of the Vu/f strut parallel to the beam span (Figure 2.8c), the tie force
may be resisted by reinforcing steel located in the girder bulb above the bearing and extending
approximately one-quarter the girder depth into the span. That is, tie reinforcement located
within H/4 + Lbearing from the back of the bearing resists the tie force, where Lbearing is the length
of the bearing.

2.5.3  Illustrative Examples


For girder shapes having the same bulb geometry (e.g., BT and NU), the deepest such girder
will be critical because deeper girders will have the greatest design shear (being the longest
and heaviest girder). On this basis, BT-72, NU-1800, and AASHTO VI sections, all having the
same nominal depth, d = 72 in., were selected for comparison. The approach taken to develop
representative examples was as follows:
1. For each girder shape, three straight strand geometries (Nw) were selected as follows:
a. Relatively lightly reinforced (strands occupying only lowest rows in the bulb),
b. Moderately reinforced (strands populating nearly all rows in the bulb), and
c. Heavily reinforced (at least five rows of strands extending into web).
2. For each BT-72 and AASHTO VI girder, a representative length (L) was selected using the pre-
liminary design charts of Chapter 6 of the PCI Bridge Design Manual (2011). For NU-1800
girders, preliminary design charts provided in Hanna et al. (2010) were used. Values for girders
with 0.6-in. diameter strands; f c′ = 12 ksi for BT-72 and AASHTO VI and 10 ksi for NU-1800;
and girder spacings of 6 ft and 12 ft were used except as noted.
3. The design shear force, Vu/f, is calculated based on AASHTO LRFD Bridge Design Specifications
(2014); the following basic assumptions were made:
a. Girder self-weight is based on published values from the PCI Bridge Design Manual (2011)
and Hanna et al. (2010).
b. Slab self-weight is calculated using a unit weight of 0.15 kips/ft3. Slab depth was taken as
8 in. and 10 in. for the cases having girder spacing of 6 ft and 12 ft, respectively.
c. An additional 0.3 kips/ft dead load is applied to each girder accounting for other bridge
components (1.5 kips/ft on bridge distributed over five girders).
d. AASHTO HL-93 live load arranged to maximize shear at the girder critical section for shear.
e. Distribution factors for shear are calculated assuming interior girder having multiple lanes
loaded (AASHTO LRFD Table 4.6.2.2.2.3a-1).
f. STRENGTH I load combination including impact factor of 1.33.
The cases considered are summarized in Table 2.7 and shown in Figure 2.10. All cases
use straight 0.6-in. diameter strand, and a bearing width 2 in. less than the bulb soffit width
(i.e., bb = B2 - 2 in.). Bearing length is taken as 12 in. in all cases.
Changing these assumptions will have the following general effects:
1. Tie force varies minimally with strand diameter. For the same girder cross section, the change
in moment capacity is approximately proportional to the change in strand area, Mn ≈ Apsfpudv,
whereas the corresponding change in achievable girder length, L, varies approximately with
√Mn; thus, the design shear (which is a function of L) and, therefore, tie force will be smaller
for 0.5-in. diameter strands than for 0.6-in. diameter strands. 0.7-in. diameter strands will
similarly increase the tie force but not proportionally as both moment and span length will

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Analytical Research Approach and Findings   35  

Table 2.7.   Summary of girder geometries considered and resulting tie requirements.
Girder geometry Range of Tie forces Cases satisfied by 60 ksi ties…
H S N at L V/ Cases debonding tmax tdr = 0 No. 3 @ 6 in. No. 4 @ 6 in.
Girder
in. ft midspan ft kips ratio kips kips having capacity…
83 kips 151 kips
BT-72 82 12 24 @ 0.6 95 372 32 0 – 0.67 228 145 25% 62%
BT-72 80 6 24 @ 0.6 135 297 32 0 – 0.67 182 116 31% 88%
BT-72 82 12 38 @ 0.6 115 421 98 0 – 0.63 131 100 60% 100%
BT-72 80 6 38 @ 0.6 138 326 98 0 – 0.63 101 77 88% 100%
BT-72 82 12 48 @ 0.6 122 437 98 0 – 0.50 71 53 100% 100%
BT-72 80 6 48 @ 0.6 163 339 98 0 – 0.50 55 41 100% 100%
BT-72 82 12 38 @ 0.5 97 377 98 0 – 0.63 118 89 76% 100%
BT-72 80 6 38 @ 0.5 130 290 98 0 – 0.63 91 69 96% 100%
BT-54 64 12 38 @ 0.6 94 364 98 0 – 0.63 114 86 65%a 100%a
BT-54 62 6 38 @ 0.6 130 281 98 0 – 0.63 88 67 88%a 100%a
a
For BT-54, capacity of No. 3 ties at 6 in. = 71 kips and No. 4 ties at 6 in. = 130 kips.
AASHTO VI 82 12 24 @ 0.6 90 379 21 0 – 0.58 58 49 100% 100%
AASHTO VI 80 6 24 @ 0.6 131 318 21 0 – 0.58 49 41 100% 100%
AASHTO VI 82 12 48 @ 0.6 128 478 156 0 – 0.63 81 77 100% 100%
AASHTO VI 80 6 48 @ 0.6 165 376 156 0 – 0.63 63 60 100% 100%
AASHTO VI 82 12 76 @ 0.6 145 521 633 0 – 0.55 63 63 100% 100%
AASHTO VI 80 6 76 @ 0.6 180 401 633 0 – 0.55 49 48 100% 100%
NU-1800 81 12 24 @ 0.6 88 360 36 0 – 0.67 327 120 28% 58%
NU-1800 79 6 24 @ 0.6 120 282 36 0 – 0.67 256 94 36% 75%
NU-1800 81 12 48 @ 0.6 136 478 583 0 – 0.67 390 220 13% 37%
NU-1800 79 6 48 @ 0.6 167 355 583 0 – 0.67 289 164 22% 63%
NU-1800 81 12 60 @ 0.6 153 519 1165 0 – 0.60 215 145 61% 96%
NU-1800 79 6 60 @ 0.6 186 384 1165 0 – 0.60 159 107 77% 100%
NU-1100 53 12 24 @ 0.6 73 313 36 0 – 0.67 283 104 19% 39%
NU-1100 51 6 24 @ 0.6 90 224 36 0 – 0.67 203 75 31% 64%
NU-1100 53 12 48 @ 0.6 105 392 583 0 – 0.67 319 181 11% 30%
NU-1100 51 6 48 @ 0.6 128 281 583 0 – 0.67 229 129 19% 52%
NU-1100 53 12 60 @ 0.6 115 415 1165 0 – 0.60 172 116 56% 87%
NU-1100 51 6 60 @ 0.6 142 301 1165 0 – 0.60 125 84 72% 99%
b
For NU-1100, capacity of No. 3 ties at 6 in. = 59 kips and No. 4 ties at 6 in. = 108 kips.

also be limited by other design constraints (e.g., release stresses). Figure 2.11 illustrates a
comparison of tie forces calculated for a BT-72 girder having 38 0.6- or 0.5-in. diameter
strands.
2. Reducing the girder depth will reduce both the moment capacity and achievable span length.
The design shear is approximately proportionally to span; thus, shallower girders, while hav-
ing a greater shear to moment ratio, will, nonetheless, have a lower design shear and, hence,
lower tie force. Figure 2.12 illustrates a comparison of tie forces calculated for BT-72 and
BT-54 girders having 38 0.6-in. diameter strands.
3. Changing f c′ has no impact on the present calculations but would affect the maximum shear
that may be resisted by a section, given by AASHTO LRFD Article 5.8.3.3 as Vn ≤ 0.25f c′ bvdv.
This value would only be reached for heavily prestressed girders having impractically short
spans.
4. Decreasing the bearing width, bb, increases the tie force considerably. Full width bearing is
recommended for single-web bulbed girders. Full width is understood to mean the full width
of the bottom flange less a distance accounting for the chamfer—typically 2 in. on both sides.
Figure 2.13 illustrates the effect of reduced bearing width for BT-72 girders.
5. Bearing length has no impact on tie force but a shorter bearing length reduces the length of
the region over which the resisting ties need to be placed (i.e., H/4 + Lbearing).
A total of 9,408 cases were developed by investigating four girder shapes (BT-72, AASHTO VI,
NU-1800, and NU-1100) in combination with two values of girder spacing (6 ft and 12 ft). For

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

36   Strand Debonding for Pretensioned Girders

6 in. 6 in. 6 in.

72 in. + slab

72 in. + slab

72 in. + slab
10.5 in.

10.5 in.

10.5 in.
24 in. 24 in. 24 in.
26 in. 26 in. 26 in.
BT-72 BT-72 BT-72
N = 24 N = 38 N = 48

250

200

150 #4 ties @ 6 in.


Tie force, t (kips)

fy = 60 ksi

100

#3 ties @ 6 in.

50

compressive strut
predicted
-50
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
Debonding ratio, dr
(a) BT-72

Figure 2.10.   Tie forces.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Analytical Research Approach and Findings   37  

8 in. 8 in. 8 in.

72 in. + slab
72 in. + slab

72 in. + slab
18 in.

18 in.

18 in.
26 in. 26 in. 26 in.
28 in. 28 in. 28 in.
AASHTO Type VI AASHTO Type VI AASHTO Type VI
N = 24 N = 48 N = 76

100

#3 ties @ 6 in.

75 fy = 60 ksi

50
Tie force, t (kips)

25

compressive strut
predicted
-25

-50
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
Debonding ratio, dr
(b) AASHTO Type VI

Figure 2.10.  (Continued).

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

38   Strand Debonding for Pretensioned Girders

5.9 in. 5.9 in. 5.9 in.

72 in. + slab

72 in. + slab

72 in. + slab
10.8 in.

10.8 in.

10.8 in.
36.4 in. 36.4 in. 36.4 in.
38.4 in. 38.4 in. 38.4 in.

NU-1800 NU-1800 NU-1800


N = 24 N = 48 N = 60

350

300

#4 ties @ 3 in.
250

200
Tie force, t (kips)

150 #4 ties @ 6 in.

fy = 60 ksi
100
#3 ties @ 6 in.

50

compressive strut
-50 predicted
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
Debonding ratio, dr
(c) NU-1800

Figure 2.10.  (Continued).

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Analytical Research Approach and Findings   39  

5.9 in. 5.9 in. 5.9 in.

72 in. + slab

72 in. + slab

72 in. + slab
10.8 in.

10.8 in.

10.8 in.
36.4 in. 36.4 in. 36.4 in.
38.4 in. 38.4 in. 38.4 in.

NU-1100 NU-1100 NU-1100


N = 24 N = 48 N = 60

350

300

250

#4 ties @ 3 in.
200
Tie force, t (kips)

150

#4 ties @ 6 in.
100
fy = 60 ksi

#3 ties @ 6 in.
50

compressive strut
-50 predicted
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
Debonding ratio, dr
(d) NU-1100

Figure 2.10.  (Continued).

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

40   Strand Debonding for Pretensioned Girders

250

200

150 #4 ties @ 6 in.


Tie force, t (kips)

fy = 60 ksi

100

#3 ties @ 6 in.

50

compressive strut
predicted
-50
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
Debonding ratio, dr

Figure 2.11.   Tie forces for BT-72 having N = 38 0.6- and 0.5-in. diameter strands
and spans appropriate for girder spacing of 12 ft.

250

200

#4 ties @ 6 in.
150 BT-72

BT-54
Tie force, t (kips)

fy = 60 ksi
100
#3 ties @ 6 in.
BT-72
BT-54

50

compressive strut
predicted
-50
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
Debonding ratio, dr

Figure 2.12.   Tie forces for BT-72 and BT-54 having N = 38 0.6-in. diameter strands
and spans appropriate for girder spacing of 12 ft.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Analytical Research Approach and Findings   41  

250

200

150 #4 ties @ 6 in.


Tie force, t (kips)

fy = 60 ksi

100

#3 ties @ 6 in.

50

full-width
bearing pad
bb = 24 in.
0

compressive strut
predicted
-50
10 12 14 16 18 20 22 24 26
Bearing width, bb (in.)

Figure 2.13.   Effect of variation of bearing pad width, bb , on BT-72 girders having
dr = 0.0.

each girder/spacing combination, all possible debonding patterns of strands in the outer portion
of the bulb (i.e., varied nf) that satisfied the following criteria were considered:
1. Outermost strand in lowermost full-width rows remain bonded;
2. No more than 50% debonding in first (lowest) row; and
3. No strands in the plane of the web (i.e., the shaded region in Figure 2.9) were debonded.
Figure 2.10 shows the calculated tie forces, t, for all cases plotted against the debonding ratio, dr.
Due to the interaction of xp, yp, and cb associated with debonding patterns, there is no trend in tie
force although more debonded cases tend to have tie forces less than those calculated for the fully
bonded case (dr = 0) than greater. Also shown in Figure 2.10 is the tie capacity of 60 ksi No. 3
and No. 4 ties (i.e., two legs) having a spacing of 6 in. distributed over the length H/4 + Lbearing.
Tie capacity may be proportionally increased using ties with strengths of 75 or 100 ksi, and
capacities at other spacing may be similarly interpolated. Using “bundled” ties is also a means
of increasing capacity. For example, two bundled No. 3 ties have 10% greater capacity than one
No. 4 tie. Similarly, bundled ties at 6-in. spacing may be more practical to install and result in less
congestion than single ties at 3-in. spacing, for instance. Additionally, embedded steel sole plates
may contribute significantly to the available tie force. Bearings that are very stiff in plane (steel
plate) may also contribute to the tie force through the force that may be transmitted through
friction with the concrete soffit. Typical neoprene bearings are not believed to be sufficiently stiff
to contribute to the tie force in a meaningful manner. Table 2.7 summarizes the maximum tie
force observed for each beam detail, and the number of cases satisfied using each tie geometry
by assuming the ties alone provide the required force.
The following conclusions are drawn:
1. Tie force is dependent on the girder shape. Assuming a fixed value of bearing width, bb, tie
force, t, is most affected by dimension xp as it defines the slope of strut A (Figure 2.9). As such,

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

42   Strand Debonding for Pretensioned Girders

tie forces are greater for wide flat bulbs (NU-1800) than for thinner deep bulbs (AASHTO
Type VI). BT-72 falls between these extremes. A rectangular stem beam would have no tie force
(that is, nf = 0).
2. Tie forces are reduced by maximizing the middle strut proportion of shear, i.e., (Vu/f)
(1 - 2nf /Nw). The large AASHTO Type IV through VI girders accomplish this objective with
a wider web resulting in four, rather than two, vertical columns of strands contributing to the
middle strut. This observation additionally reinforces the guidance not to debond strands
within the plane of the web.
3. No simple “rule of thumb” for detailing is identified. It is concluded that No. 3 ties at 6-in.
spacing are adequate for AASHTO type girders. Similarly, No. 4 ties at 6-in. spacing are adequate
for all but lightly prestressed BT-72 sections.
4. Due to their bulb width, deep NU girders have high tie forces and, therefore, require con­
siderable more tie reinforcement (greater than No. 4 ties at 3 in. in many cases). To investigate
the NU shape further, the analysis was repeated for a shallower NU-1100 section. The results
are shown in Figure 2.10d and summarized in Table 2.7. Although the tie forces are reduced, the
length over which the ties are provided, (H/4 + Lbearing), is also reduced; thus, tie reinforcement
requirement is affected marginally.

2.5.4  Validation by Experimental Results


The STM was applied to both the design and as-built experimental behavior of the single-web
bulbed beams tested as part of this study. These results are presented in Section 3.5.
The approach was also applied to the Type IV girder A2U1 tested by Hamilton (2009) that
exhibited clear evidence of longitudinal splitting (Figure 2.8). The results are shown in Table 2.8.
Although no details or photographs of the bulb confinement reinforcement are available, typically
No. 3 ties at 6 in. would likely be provided, which is approximately one-half the reinforcement that
would be required to resist the tension tie developing at the ultimate girder capacity. As shown in
Figure 2.8, the girders were supported on neoprene bearings having a width of only 18 in. Little

Table 2.8.   STM of girders tested by others.


Girder 6
A2U1 BT-7-Live
(Hawkins and Kuchma
(Hamilton 2009) (Russell et al. 2003)
Variable 2007)
Type IV BT-72 BT-63
East West
dr 0.57 0.25 0.00 0.38
H (in.) 62 80 73
N 15 18 42 26
nf 6 7 14 10
xp (in.) 8.3 7.9 6.3 6.4
yp (in.) 5.0 3.6 4.0 4.8
hb (in.) 17 10.5 10.5
bb (in.) 18 24 24
cb (in.) 5.4 7.3 8.0 7.4
0.512 0.502 0.179 0.353
Vdesign. (kips) 478 582 536
t = Vdesign (kips) 240 104 189
not reported
No. 3 ties at No. 3 ties at
ties required over H/4 + Lbearing No. 3 ties at 1.6 in.
3.8 in. 2.0 in.
Vexp. (kips) 296 614 640 541
t = Vexp (kips) 152 308 115 191
No. 3 ties at 2.5 No. 3 ties at No. 3 ties at
ties required over H/4 + Lbearing No. 3 ties at 1.3 in.
in. 3.3 in. 1.9 in.
ties provided over H/4 + Lbearing No. 3 ties at 6 in. No. 3 hairpins at 6 in. No. 3 ties at 5 in.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Analytical Research Approach and Findings   43  

restraint would be expected from this bearing; thus, the observed splitting should be expected.
Increasing the bearing to full width (24 in.) would reduce the tie force by 30% although the force
would still have exceeded the capacity of No. 3 ties at 6 in. This trend is an indication that the
debonding pattern in this girder drove the observed splitting behavior.
Three BT-72 girders reported by Russell et al. (2003) investigated varying shear reinforcement
designs. Each girder had the same strand and debonding pattern, and was reportedly provided
with “No. 3 hairpins at 6 in.” as bulb confinement reinforcement. Girder BT-7-Live had the
largest design load and experimental capacity; the STM analysis of this girder is reported in
Table 2.8. Like A2U1, it appears that BT-7-Live had inadequate tie reinforcement required to
resist the tension tie developing at the ultimate girder capacity. This girder was supported on a
steel bearing plate that may be expected to provide some degree of tie restraint due to friction
between the girder and transversely stiff plate. Such bearing restraint is not accounted for in
the STM prediction. While no splitting is reported (this was not a concern of the test program
and, therefore, may have simply been neglected), significant strand slip was reported for a single
strand, which may indicate some internal transverse distress.
Girder 6, reported by Hawkins and Kuchma (2007) and described in Appendix C was also
modeled. The east end of this girder had 42 fully bonded straight strands while the west end
had 16 of these strands debonded. Once again, the provided tie reinforcement appeared to be
inadequate to resist the expected tie force developed at the ultimate limit state. However, it is
not clear from the girder details whether some of the web shear reinforcement was anchored in
such a way as to provide additional tie restraint. Additionally, like BT-7, steel bearing plates were
used which would be expected to provide additional restraint.

2.5.5  Summary of End Region Behavior


The current requirement for confinement reinforcement in AASHTO LRFD Article 5.10.10.2
(i.e., No. 3 ties having spacing not exceeding 6.0 in.) is apparently adequate to resist tension tie
forces at the ultimate limit state for sections with a narrow bottom bulb (e.g., AASHTO I girders).
However, this requirement is not sufficient for sections such as BT and NU girders that have a
wider bulb.
The proposed STM is a simple and effective method for assessing the demand on the confine-
ment reinforcement in the bottom flange, and determining the required amount of reinforce-
ment. In lieu of using the STM approach presented herein, No. 4 ties spaced at 3 in. provide
adequate capacity for all cases except NU girders with fewer than 60 strands, which is the maxi-
mum number of strands that can fit in the bottom bulb. For these cases, the STM formulation
may be used or an embedded steel sole plate is recommended. It should be noted that the cur­
rent Nebraska standard details require the use of such plates for NU girders [Jaber 2016; Bridge
Office Policies and Procedures (BOPP) 2011].

2.5.6 Extension of Bulbed Girder Results and Discussion


of Other Girder Shapes
The issues described in the previous sections are only applicable to single-web sections. In
double-web sections such as box or U, shear is transmitted through the webs to the level of the
bottom flange. Considering the idealized STM shown in Figure 2.14, there is no need for the web
compressive strut to spread to find an anchoring node if the following two conditions are met:
1. Bearing is provided under at least the full dimension of each web (accounting for the
chamfer—typically 2 in.); the full-width support allows the strut to be anchored in the same
plane as it develops.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

44   Strand Debonding for Pretensioned Girders

CL

Figure 2.14.   STM of box


girder.

2. The strands located in the planes of the webs are not debonded; this detail allows the necessary
tension force also to be developed in the same plane as the compressive strut. As a result, partial
debonding in box or U-sections should be restricted to the flange region outside the web width.
Box or U-girders may also have end diaphragms to help to spread the web-delivered shear
across the flange at the support. In such cases, small horizontal compressive struts, rather than
tension tie demands will develop.

2.6  Evaluation of the Effects of Skewed Girder Ends


Cracking at the flange-web interface of prestressed girders with skewed ends has been observed
in practice. Figure 2.15 shows an example of such cracking. The effects of skewed ends were
examined through a mechanistic model and nonlinear FEM.

2.6.1  Mechanistic Modeling


A model was developed with reference to the generic bulbed girder shown in Figure 2.16, in
which OB is the flange-web interface having a flange thickness of tf . OA is the section defining
the triangular region of flange projecting beyond the web due to the skew angle q.

Cracks
Shrinkage cracks

Figure 2.15.   Example of cracking in skewed ends


in W74G with 55-degree skew angle. Photo Courtesy
of Washington State DOT.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Analytical Research Approach and Findings   45  

Figure 2.16.   Geometry and internal force distribution at the end of a skewed
bulbed girder.

The developed length of bonded strand i at OA is (Eq. 2.10):

Li = yi tan θ Eq. 2.10

The force associated with each vertical group of ni bonded strands at location i developed at
section OA is (Eq. 2.11):

Li y tan θ
Pi − f pini Aps = f pini Aps Eq. 2.11
lt lt

The moment about OO ′ due to the strands is ( Eq. 2.12): M o = ∑ Pi yi Eq. 2.12

The shear along plane OB due to the strands is ( Eq. 2.13): V = ∑ Pi Eq. 2.13

M O develops tensile stresses ( ft ) along the flange-web interface having a peak value at point O

6∑ Pi yi  x
and assumed to be triangularly distributed over length L ( Eq. 2.14 ): ft = 1− 
t f L2  L
Eq. 2.14

V is resisted along the flange-web interface as a function of the anticipated shear lag across the
flange dimension y. Struts having an angle a as shown may be used to approximate this behavior.
In this manner, L may be defined by the choice of a, which may be reasonably estimated to be
26.6 degrees, i.e., L = 2 yi,max, which also corresponds approximately to that observed in Figure 2.15.

2∑ Pi  x 
The shear stress distribution along plane OB is (Eq. 2.15): v = Eq. 2.15
tf L  L 

ft f 2t
The interaction of ft and v results in a principal tensile stress ( Eq. 2.16): σ1 = + + v2
2 4
Eq. 2.16

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

46   Strand Debonding for Pretensioned Girders

2v
This principal stress is oriented at an angle β from OA ( Eq. 2.17 ): tan 2β = Eq. 2.17
ft

For these calculations, all terms are defined in Figure 2.16 with the exception of:
Aps = Area of one prestressing strand
fpi = Initial prestress force
lt = Strand transfer length
ni = Total number of strands in group i of strands
Based on the above equations and interdependence of the parameters selected, a number of
conclusions can be made:
1. The strand forces in the triangular region of flange projecting beyond the web (Eq. 2.11) are
proportional to the tangent of the skew angle, tan q. As a result, MO and V are also propor-
tional to tan q.
2. The strand forces in the triangular region of flange projecting beyond the web (Eq. 2.11),
and, therefore, MO and V are inversely proportional to transfer length of the strand. Thus, a
shorter transfer length results in higher forces, the corollary of which is that the use of a longer
assumed transfer length for design is unconservative in this case.
3. Debonding any strands in the triangular region of flange projecting beyond the web will
reduce Pi and, thus, reduce the stresses along the web-flange interface.
4. Coupling Observations 1 and 3 suggests that some debonding is required in highly skewed
bulbed girders in order to reduce web-flange interface stress.

2.6.2  Washington State Girder Examples


The previously described approach and the observations stemming from it were evaluated
with reference to girders G2 and G5 provided by Washington DOT. Both girders were W74G
with end skews of 55 degrees. Each girder had 0.6-in. diameter Gr. 270 strands jacked to 43.9 kips,
and had 6.0 ksi concrete. Girder G5 (details shown in Figure 2.17b) had 24 fully bonded straight
strands in the flange, and exhibited cracking believed to be associated with the skew geometry
(Figure 2.15). Girder G2 (Figure 2.17a), on the other hand, exhibited no such cracking. This
girder had the same geometry as G5 but only eight straight strands and a detail in which the two
strands at the acute corner were debonded a distance of 29 in. Key girder details are shown in
Figure 2.17.
The calculated principal stresses (Eq. 2.9) plotted in Figure 2.18 confirm that girder G5 is
expected to crack but girder G2 is not. In these calculations, a realistic in situ transfer length of
30db was assumed. Using Girder G5, having 24 straight strands as a benchmark, the amount of
debonding required to prevent cracking was calculated for different skew angles. The results
shown in Table 2.9 indicate that for skew angles less than 19.5 degrees, no debonding of the
24 strands is required. For large skew angles greater than 55 degrees, all the strands in groups 1,
2, and 3 would have to be debonded to prevent the cracking observed (it is not recommended to
debond the most exterior strands in a row). Table 2.9 also indicates that one-half of these strands
will have to be debonded for skew angles greater than 25 degrees, and all of these strands will
have to be debonded for skew angles greater than 55 degrees.

2.6.2.1  Nonlinear Finite Element Modeling of Girders G2 and G5


Nonlinear FEM (Section 2.4) was performed to evaluate the aforementioned observations,
and further examine the effects of skewed ends. Using girders G2 and G5 (Figure 2.17) as a
benchmark, five cases were examined, as summarized in Table 2.10. Cases G2-B and G5-A are

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Analytical Research Approach and Findings   47  

MARKED END MARKED END

(a) Girder G2 (note debonding of two strands (b) Girder G5


at acute corner [left side])

Figure 2.17.   Details of girders G2 and G5. Drawings: Courtesy of Concrete


Technology Corp., Tacoma, WA.

intended to replicate the field observations (hence, the use of a “realistic” transfer length of 30db,
rather than the “design” transfer length of 60db). The remaining cases are intended to examine
the following objectives:
1. The objective for G2-A was to assess whether the lack of cracking in G2-B was due to
debonding of the acute side strands or the lower overall prestress force in this girder com-
pared to that in G5-A. The acute side strands were left fully bonded in G2-A.
2. G5-B was used to assess whether the observed cracking is predicted when using the design
transfer length of 60db. In other words, is the observed cracking a result of the developed

2.5

2.0

1.5
(ksi)

1 (ksi): G2
1 (ksi): G5
1.0
ft (ksi)

0.5

0
0 0.2 0.4 0.6 0.8 1. 0
x/L

Figure 2.18.   Variation of principal tensile stresses.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

48   Strand Debonding for Pretensioned Girders

Table 2.9.   Influence of skew angle on the required debonding


for preventing cracking.

Number of Bonded Strands in One-Half of


Skew angle the Bulb for Which No Cracking Is
Predicted Total dr
(deg.)
Group 1 Group 2 Group 3 Group 4
< 19.5 3 3 2 2 0.00
20 3 2 2 2 0.09
25 3 2 1 2 0.17
30 3 2 1 1 0.25
35 3 1 1 1 0.33
40 3 1 0 1 0.42
45 3 0 0 1 0.50 4 3 2 1
50 3 0 0 1 0.50
55 3 0 0 0 0.58
60 3 0 0 0 0.58
65 3 0 0 0 0.58

force at line OA shown in Figure 2.16 being greater in the “realistic” case than in the
“design” case?
3. Based on the outcome of G5-A and G5-B, G5-C was designed to mitigate cracking, hence,
demonstrating the hypothesis presented. The detail used for G5-C repeated the analysis of
G5-A (30db) with six straight strands debonded as shown in Figure 2.19.
A summary of predictions of cracking at the time of prestress release from skewed girder
modeling is shown in Figure 2.20. Since minor cracking was observed in G5-A and not G5-B,
the “difference” between realistic (30db) and design (60db) transfer length, alone, is apparently
sufficient to account for cracking in this girder regardless of the large prestress force. For G5-C,
the objective was to provide sufficient debonding to mitigate the observed cracking using the
realistic 30db transfer length. It was found that debonding 6 strands (indicated in Figure 2.19) for
a length of 29 in. was sufficient to mitigate observed cracking when using a transfer length of 30db.
This observation was confirmed by also considering 4 strands debonded (cracking was observed)
and 8 strands (no cracking); the latter cases are not shown.
Based on the results summarized in Figure 2.20, the following observations are made:
1. The analyses confirm the research team’s hypothesis that no cracking would be observed in
G2-A or B. Thus, no conclusion may be drawn on the effects of prestressing or transfer length.
The transverse moment generated (Eq. 2.12) with only 8 strands in G2 is simply too small to
result in cracking.

Table 2.10.   Summary of skewed girder FEM analyses.


Case G2-A G2-B G5-A G5-B G5-C
Straight strands 8 8 24 24 24
6 strands
Debonded
None 2 at acute corner None None debonded at
strands
29 in.
Transfer length
30db 30db 30db 60db 30db
used in FEM
Field Cracking (see
N.A. No cracking N.A. N.A.
observations Figure 2.15)
No cracking at any No cracking Minor cracking No cracking No cracking
threshold (Fig. predicted at any predicted at predicted at predicted at
FEM-predicted
2.20a) threshold 0.008 in. 0.008 in. 0.008 in.
cracking
(Fig. 2.20b) threshold (Fig. threshold threshold
2.20c) (Fig. 2.20d) (Fig. 2.20e)

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Analytical Research Approach and Findings   49  

Debonded
strands

(a) Cross section (b) FEM

Figure 2.19.   Strand debonding in girder G5-C (debonded strands


are circled).

(a) Girder G2-A:


8 straight strands,
no debonding,
30db transfer
length. No
cracking
predicted
(b) Girder G2-B:
8 straight strands,
2 debonded
strands, 30db
transfer length.
No cracking
predicted

(c) Girder G5-A:


24 straight
strands, no
debonding, 30db
transfer length.
Cracking
predicted
(d) Girder G5-B:
24 straight
strands, no
debonding, 60db
transfer length.
No cracking
predicted

(e) Girder G5-C:


24 straight
strands, 6 strands
debonded, 30db
transfer length.
No cracking
predicted

Figure 2.20.   Crack predictions at 0.008 in. threshold from


skewed girder modeling (only predicted cracks greater than
0.008 in. are shown).

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

50   Strand Debonding for Pretensioned Girders

2. G5-A exhibited cracking very similar to that seen in the field (Figure 2.15).
3. G5-B exhibited essentially no cracking. Hence, the “more gradual introduction of prestress
force” resulting from longer “design” transfer length of 60db is sufficient to mitigate the crack-
ing observed. That is, the individual strand forces, Pi (Eq. 2.11) and the transverse moment,
MO (Eq. 2.12) is reduced, thereby lowering the concrete stress sufficiently to mitigate the
observed cracking.
4. G5-C accomplishes the same objective as G5-B but through debonding in order to reduce the
transverse moment, MO.
The following conclusions are drawn:
1. The flange-web cracking shown in Figure 2.15 is attributable to transverse moment as described.
Cracking is more significant with a greater moment present, and limiting this moment may
mitigate cracking.
2. Reducing the nonuniform prestress force at a transverse section (i.e., OA in Figure 2.16) may
mitigate the transverse moment in a skewed-ended girder. This goal may be accomplished
by limiting the prestress force itself (compare G5-A to G2-A) or by providing debonding to
minimize the moment at the root of the web at plane OO′ (compare G5-C to G5-A).
3. The use of more realistic transfer lengths (i.e., those shorter than the AASHTO-prescribed 60db)
is more appropriate when evaluating the potential cracking of skewed ends. The longer design
values result in lower stresses and are, therefore, non-conservative with respect to predicting
the type of cracking shown in Figure 2.15.

2.7  Introduction of Debonded Strands at One Section


The test specimens (Chapter 3) were detailed such that the debonded strands were gradually
introduced along the girder length. This goal was achieved by meeting the following criterion
stipulated in AASHTO LRFD Section 5.11.4.3: “Not more than 40 percent of the debonded strands,
or four strands, whichever is greater, shall have the debonding terminated at any sections.” Similarly,
the debonded strands in all the 44 cases described in Section 2.4.2 and Appendix D were rebonded
at intervals of 3 ft.
To investigate the case in which a large number of debonded strands are bonded at a single
section, Case 13B was introduced. Case 13 (see Table 2.3) is an NU-900 having 28 fully bonded
strands, 12 strands bonded at 3 ft, and 10 additional strands bonded at each of 6 ft and 9 ft, making
a total number of strands N = 60 and maximum debonding ratio is dr = 0.53. Case 13B is otherwise
identical to Case 13 except all 32 debonded strands are bonded at the same section at 6 ft.
As shown in Table 2.6, this concentrated debonding pattern had little effect on the stresses at
prestress release, resulting in a 4.5% increase in maximum compressive stress and 16% decrease
in tensile stress. Figure 2.21 shows the stress distributions at prestress release. The effect of the
sudden introduction of 32 strands at 6 ft (Case 13B), rather than being introduced in three
increments over 9 ft (Case 13) is evident as a more abrupt transition of stress (marked with an
arrow in Figure 2.21). While this stress raiser does not appear to result in any over-stresses, it
should nonetheless be avoided by distributing the debonding curtailment in a uniform manner
as presently required by AASHTO.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Analytical Research Approach and Findings   51  

Elevation - Longitudinal stress (Vertical dimension exaggerated 200%)


Bearing -3.91 ksi Bearing -3.91 ksi

-2.61 ksi -3.26 ksi -2.61 ksi -3.26 ksi


Reverse plan (soffit) - Longitudinal stress (Transverse dimension exaggerated 200%)

Reverse Plan (soffit) - Transverse stress (Transverse dimension exaggerated 200%)


all stresses shown are compressive stress less than 0.15 ksi

(a) Case 13 (b) Case 13B

Figure 2.21.   Effect of gradual versus concentrated introduction


of debonded strands on stresses at prestress transfer. (Only the
initial 25 ft of 100 ft long girder shown in all images.)

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Chapter 3

Experimental Research Approach,


Findings, and Associated
Analytical Simulations
3.1  Research Approach
The effects of strand debonding on the performance of prestressed girders were examined
experimentally through design, fabrication, and testing of six girders. The data obtained at pre-
stress release and at various stages of load testing were utilized for this purpose. The FEM plat-
form described in Section 2.4 and STM discussed in Section 2.5 were utilized to develop a better
understanding of the experimental data.

3.2  Design and Detailing of Test Specimens


Both ends of six full-scale girders were tested for a total of 12 sets of results. The main test vari-
ables were (1) girder shape (single-web, box, or U), (2) amount of debonding of partially debonded
strands, (3) concrete strength, and (4) strand diameter. Each girder had different debonding ratios
at ends A and B. Key aspects of the specimens are summarized in Table 3.1. The specimen details
are shown in Appendix E. With the exception of AASHTO BI-36, all the girders had a 6-in. thick
slab over the entire width of the top flange. The deck slab reinforcement was designed according
to the empirical design procedure described in AASHTO LRFD Article 9.7.2.5. No slab was pro-
vided on the BI-36 girder; additionally, a 2.5-ft. thick end diaphragm was provided to replicate
a common practice used in box girders.
The test girders were designed and detailed to satisfy the following criteria:
a. Satisfy concrete tensile stress limits at prestress release (AASHTO LRFD Article 5.9.4.1.2).
b. Check AASHTO LRFD Eq. 5.8.3.5-2 from the interior face of support to the critical section
to ensure that there is adequate Aps fps, accounting for debonding.
As debonded strands become bonded, their pretensioning force is gradually developed.
Figure 3.1 compares transfer of pretensioning forces in a girder with bonded and partially
debonded strands. In this representative case, the full capacity of all strands in the section,
Aps fps, is not available before 21 ft into the span. The calculations were based on the current
AASHTO-prescribed equation for strand development length:

 2 
ld = κ  f ps − f pe  db [5.11.4.2.1]
 3 

The depths of the test girders were greater than 24 in.; hence, the value of k was taken as 1.6
for fully bonded strands. (The value of k would have been 1.0 had the girders been shallower
than 24 in.) For the partially debonded strands, a value of 2.0 was used for k.

52

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Experimental Research Approach, Findings, and Associated Analytical Simulations   53  

Table 3.1.   Key details of the test specimens.

Debonding Ratio Reinforcement


Total
db Total Longitudinal Transverse (No. 4)
Girder End No. of 0 to 3 to 6 to 9 to
(in.) Per >12' No. & Cutoff Web
Strands 3' 6' 9' 12' Bottom flange
Row Size Point (ft) U shaped
Section 0.50 0.36 0.18 0.09 0
Row 2 spaces @ 3 in.
A Row 1 0.40 0.27 0.13 0 0 4 No. 6 8.5 @ 12 in.
1: 15
AASHTO Row 2 0.71 0.57 0.29 0.29 0 (outside of
0.5 7 spaces @ 6 in.
BI-36 Section 0.18 0.09 0 0 0 end
Row diaphragm) @ 12 in. to
B Row 1 0.13 0 0 0 0 None N/A
2: 7
Row 2 0.29 0.29 0 0 0 midspan
Section 0.60 0.40 0.20 0 0 2 No. 6 6.5
Row 4 spaces @ 8 spaces @ 3 in.
A Row 1 0.40 0.40 0 0 0
1: 10 8 No. 6 13.5 3 in.
AASHTO Row 2 0.80 0.40 0.40 0 0 10 spaces @ 6
0.6
BT-54 Section 0.10 0 0 0 0 in.
Row @ 18 in. to
B Row 1 0.20 0 0 0 0 6 No. 6 6.5 @ 18 in. to
2: 10 midspan
Row 2 0.00 0 0 0 0 midspan
Section 0.50 0.25 0.13 0 0 2 No. 5 5.5
Row 3 spaces @ 3 spaces @ 3 in.
A Row 1 0.25 0 0 0 0
1: 8 6 No. 5 10.5 3 in.
AASHTO Row 2 0.75 0.50 0.25 0 0 10 spaces @ 6
0.5
Type III-a Section 0.25 0.13 0 0 0 2 No. 5 6.5 in.
Row @ 18 in. to
B Row 1 0.25 0 0 0 0 @ 18 in. to
2: 8 4 No. 5 9.5 midspan
Row 2 0.25 0.25 0 0 0 midspan
Section 0.56 0.33 0.11 0 0 2 No. 5 5.5 3 spaces
Row 3 spaces @ 3 in.
A Row 1 0.25 0 0 0 0
1: 8 6 No. 5 10.5 @ 3 in.
AASHTO Row 2 0.80 0.60 0.20 0 0 10 spaces
0.5
Type III-b Section 0.22 0.11 0 0 0 4 No. 5 5.5 @ 18 in. @ 6 in.
Row
B Row 1 0.25 0 0 0 0 to @ 18 in. to
2: 10 4 No. 5 9.5
Row 2 0.20 0.20 0 0 0 midspan midspan
Section 0.45 0.27 0.18 0.09 0 3 spaces
Row 3 spaces @ 3 in.
A Row 1 0.44 0.33 0.22 0.11 0 6 No. 6 5.5
1: 18 @ 3 in.
Nebraska Row 2 0.50 0 0 0 0 10 spaces
0.7
NU-1100 Section 0.27 0.18 0.18 0.18 0 @ 12 in. @ 6 in.
Row
B Row 1 0.33 0.22 0.22 0.22 0 4 No. 6 5.5 to @ 12 in. to
2: 4
Row 2 0.00 0 0 0 0 midspan midspan
Section 0.50 0.35 0.19 0 0 3 spaces 3 spaces
Row 22 No. @ 3 in.
A Row 1 0.42 0.21 0 0 0 14.5 @ 3 in.
1: 19 6
Row 2 0.71 0.71 0.71 0 0 22
22 spaces
Texas spaces
0.6 Section 0.23 0.15 0.08 0 0 @ 4 in.
U-40 @ 4 in.
Row 16 No.
B Row 1 0.21 0.11 0 0 0 13.5 @ 6 in.
2: 7 6 @ 6 in. to
to
Row 2 0.29 0.29 0.29 0 0 midspan
midspan
Note: db = strand diameter.

c. Check AASHTO LRFD Eq. 5.8.3.5-1 elsewhere along span to ensure that there is adequate
Aps fps, accounting for debonding.
d. Provide longitudinal nonprestressed reinforcement (i.e., added As fy) per AASHTO LRFD
Article 5.8.3.5 if the checks in b or c are not satisfied.
e. Follow the detailing rules summarized in Table 3.2.

3.3  Material Properties


The experimentally determined concrete strengths (AASHTO Method T22) and material
properties for all the reinforcing bars (AASHTO Method T244) are summarized in Tables 3.3
and 3.4, respectively. The concrete mix designs are provided in Appendix F. All seven-wire pre-
stressing strand used was 270 ksi low-relaxation strand.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

54   Strand Debonding for Pretensioned Girders

Figure 3.1.   Development of pretensioning force in girders with partially debonded


strands.

3.4  Transfer Length


During fabrication of the girders, five vibrating wire strain gages were placed in the concrete
near the centroid of the strands at each end of each girder. The strains measured by these gages
were used to assess the in situ transfer lengths, which were compared with computed values. The
strain at each point along the girder length was determined by dividing the calculated stress by
the calculated modulus of elasticity of concrete at prestress transfer. The concrete compressive
stresses at the elevation of the vibrating wire gages (compression is negative) were calculated
based on both the gross section (Eq. 3.1) and transformed section properties (Eq. 3.2).
P Pe g ( yc, g − y ) M sw ( yc, g − y )
fc, g = − − + Eq. 3.1
Ae, g Ig Ig

Table 3.2.    Detailing rules used for the test girders.

AASHTO BT-54, AASHTO Type III, and Nebraska NU-1100


o Do not debond more than 50% of the bottom row strands.
o Keep the outermost strands in all rows located within the full-width section of the
flange (shaded region) bonded. Full width is understood to mean the full width of the
bottom flange less a distance accounting for the chamfer—typically 2 in. on both
sides.

o With the exception of the outermost strands, debond strands further from the section
vertical centerline preferentially to those nearer the centerline.
AASHTO BI-36 and Texas U-40
o Do not debond more than 50% of the bottom row strands.
o Keep the strands located in the planes of the webs bonded.
All girders
o Follow AASHTO LRFD Article 5.11.4.3: Not more than 40 percent of the debonded
strands, or four strands, whichever is greater, shall have the debonding terminated at
any section.
o Provide splitting resistance according to AASHTO LRFD Article 5.10.10.1.
o Provide confinement reinforcement according to AASHTO LRFD Article 5.10.10.2.
o Satisfy the requirements proposed in Section 4.2 of the report.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Experimental Research Approach, Findings, and Associated Analytical Simulations   55  

Table 3.3.    Measured concrete strengths (ksi).

Slab at
Age at Test
Girder End At Release, f’ci f’c at Test Time of
(days)
Test
A 102 12.6
AASHTO BI-36 7.4 N/A
B 97 12.2
A 42 17.4 11.4*
AASHTO BT-54 10.2
B 18 15.2 11.2*
A 93 12.6 7.4
AASHTO Type III-a 6.9
B 78 12.2 6.2
A 184 13.8 6.1
AASHTO Type III-b 8.3
B 155 13.2 5.7
A 67 14.0 6.9
Nebraska NU-1100 8.4
B 41 13.2 6.1
A 110 12.8 5.9
Texas U-40 6.9
B 95 12.0 5.8
*Due to scheduling issues, it was necessary to achieve at least 6 ksi in 7 days. Therefore,
the deck slab was cast using a mix design that is typically used for prestressed girders.

P Petransformed ( yc,transformed − y ) M sw ( yc,transformed − y )


fc,transformed = − − + Eq. 3.2
Ae,transformed I transformed I transformed

As shown schematically in Figure 3.2, P is introduced in the girder bulb but it is some distance
into the girder, xf , before the entire cross-sectional area is engaged in resisting P. That is, Ae,g =
Abulb or Ae,transformed = Abulb,transformed at x = 0, and Ae,g = Ag and Ae,transformed = Atransformed at x = xf . Distance
xf is defined by the assumed load-spreading angle, described by q in Figure 3.2. Angle q is taken
as 30°, which approximately corresponds to a 2:1 strut angle typically assumed in D regions. In
the transformed section calculations, the staggered bonding of the prestressing strand and the
presence of the nonprestressed reinforcement were taken into account. Moreover, “voids” in the
section representing unbonded strands were incorporated.
The effective prestressing force (P) was computed by accounting for elastic shortening deter-
mined based on AASHTO LRFD Article 5.9.5.2.3a-1 evaluated at each section. The transfer

Table 3.4.    Measured material properties


of reinforcing bars.

Girder Bar Size fy (ksi) fu (ksi) u


No. 3 82.1 120 0.126
AASHTO BI-36 No. 4 72.7 112 0.128
No. 6 65.4 102 0.186
No. 4 69.7 107 0.127
AASHTO BT-54
No. 6 65.9 106 0.132
No. 4 63.6 100 0.191
AASHTO Type III-a
No. 5 75.6 113 0.159
No. 4 63.6 100 0.191
AASHTO Type III-b
No. 5 75.6 113 0.159
No. 3 75.1 101 0.238
No. 4 79.0 106 0.254
Nebraska NU-1100
No. 5 70.1 103 0.128
No. 6 69.2 109 0.120
No. 4 70.5 110 0.157
Texas U-40 No. 5 67.1 105 0.093
No. 6 67.6 110 0.145

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

56   Strand Debonding for Pretensioned Girders

Figure 3.2.   Transition of cross-sectional area from the bulb area to


the girder area.

length was taken either as 60db (db = strand diameter) per AASHTO LRFD Bridge Design Speci-
fications or the value obtained from Eq. 3.3 (NCHRP Report 603: Ramirez and Russell 2008).
In this equation, f ci′ is the concrete strength at release, summarized in Table 3.3. The resulting
calculated transfer lengths are provided in Table 3.5.

120db
lt = ≥ 40db Eq. 3.3
f ci′

The stress determined from Eq. 3.1 or Eq. 3.2 was divided by the concrete modulus of elastic-
ity at prestress transfer (Eci) to obtain the predicted concrete strain (ec); these strains are com-
pared with the measured values. The value of Eci was determined from Eq. 3.4, which has been
published in the 2015 edition of AASHTO LRFD Bridge Design Specifications.

Ec = 120,000 K 1w c2.0 f c′ 0.33 Eq. 3.4

The value of K1 was taken as 1.0 for all the girders with the exception of Nebraska NU-1100 for
which 0.85 was used (Larson et al. 2009). The value of f c′ was set equal to f ci′ (concrete strength at
release), and 0.145 kcf was used for wc.
The calculated and measured strains are compared in Figure 3.3. Following common practice,
the effects of the end diaphragms (blocks) in girder BI-36 were neglected. The gross section prop-
erties (area and moment of inertia) are smaller than their counterparts computed from trans-
formed section properties. Hence, the strains based on gross section properties are, expectedly,
higher than those calculated from transformed section properties. With the exception of the Texas
U-40 girder, the trend of the measured data is captured reasonably well. Finite element modeling
will be presented in Section 3.7.2.2 to interpret the data for Texas U-40. At some locations and

Table 3.5.    Release concrete strengths and calculated


transfer lengths.

Calculated Transfer Lengths (in.)


Girder f’ci (ksi) db (in.)
AASHTO (60db) NCHRP 603 (Eq. 3.3)
AASHTO BI-36 7.4 0.5 30 22 (44db)
AASHTO BT-54 10.2 0.6 36 24 (40db)
AASHTO Type III-a 6.9 0.5 30 23 (46db)
AASHTO Type III-b 8.3 0.5 30 21 (42db)
Nebraska NU-1100 8.4 0.7 42 29 (41db)
Texas U-40 6.9 0.6 36 27 (45db)

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Experimental Research Approach, Findings, and Associated Analytical Simulations   57  

(a) AASHTO BI-36

(b) AASHTO BT-54

(c) AASHTO Type III-a

Figure 3.3.   Measured and computed longitudinal concrete strains at soffit.


(continued on next page)

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

58   Strand Debonding for Pretensioned Girders

(d) AASHTO Type III-b

(e) Nebraska NU-1100

(f) Texas U-40 (see Section 3.7.2.2 for additional evaluation)

Figure 3.3.  (Continued).

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Experimental Research Approach, Findings, and Associated Analytical Simulations   59  

for some of the girders, the calculations based on the current AASHTO transfer length cor-
relate better with experimental data, whereas the transfer length recommended in NCHRP
Report 603 (Ramirez and Russell 2008) yields more accurate results for some other cases. No
clear conclusion can, therefore, be made about the accuracy of either transfer length calcula-
tion method. For the majority of cases, the measured strains are larger than the computed
values. The friction between the form and girder, which is not reflected in the calculations,
would affect the boundary conditions, and, hence, the level of compressive strain in the con-
crete to a small degree.

3.5  Testing Program


The two ends of each girder (End A and End B), which had different amounts of strand
debonding, were tested separately. In each case End B was tested first. The test specimens were
extensively instrumented to capture key behavior. The test setups and instrumentation are sum-
marized in this section.

3.5.1  Test Setup


The girders were supported on neoprene pads similar to those typically used in construction.
For single-web girders, full-flange width pads, having a thickness of 1.375 in. were provided for
all the girders except for BT-54, in which 22 in. of 24.5 in. bottom flange width was supported.
For AASHTO BI-36, two 9-in. wide by 3-in. thick neoprene pads were placed under each web.
Two 3-in. thick neoprene pads were also placed under each web of Texas U-40. These pads
engaged the outer 7-in. width of the bottom flange. During testing of End B of Texas U-40, a
third pad had inadvertently been installed at the middle of the soffit at both ends. After a total
load of 120 kips, the girder was unloaded, the middle pads removed, and testing resumed. Only
two pads, one under each web, were placed at each end when End A was tested. The data pre-
sented for the Texas U-40 herein are for the second loading of End B with only the two outer
pads in place. The lengths of the pads under the soffit were 12 in. for all the girders. Displacement
transducers were attached to the girder to measure the compression of pads during testing so the
beam deflections could be corrected for this movement.
A single concentrated load was used to test the girders. The location of the load was selected
such that the shear span-to-depth ratio (a/dv) would not be less than 2.0 in order to prevent
direct transfer of the load to the support through arching action. The values of a/dv are sum-
marized in Table 3.6.
Each end was tested separately with End B tested first. After testing End B, the girder was
repositioned in order to test End A, which had a larger debonding ratio than End B. With the
exception of Texas U-40, which was tested as a simply supported span, testing of each end con-
sisted of a simple span with a propped cantilever overhang, as shown in Figure 3.4. To prevent
cracking due to the self-weight of the cantilevered portion, an air jack was used to prop the end of Table 3.6.    Shear spans
the girder. The air pressure was calibrated such that the force in the jack actively compensated for and shear span-to-
the self-weight of the cantilevered portion throughout the duration of the test; thus, the girder depth ratios.
was effectively tested as a simply supported span.
Girder a (ft) a/dv
At the conclusion of testing End B of AASHTO BT-54, a number of minor cracks were AASHTO BI-36 5.00 2.73
found to have extended into the span of End A (see Figure 3.5). In order to mitigate the effects AASHTO BT-54 10.0 2.34
AASHTO Type III-a 7.75 2.15
of these cracks, the girder was vertically post-tensioned as shown in Figure 3.6 before testing AASHTO Type III-b 7.75 2.16
End A. The total applied post-tensioning force of 120 kips was sufficient to close the small Nebraska NU-1100 7.75 2.20
cracks. Texas U-40 7.75 2.39

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

End B
Air jack
10' 6" 24' 5'
(i) Testing of End B

End A
Air jack
5' 24' 10' 6"
(ii) Testing of End A

(a) AASHTO BI-36


P

End B

Air jack 10'


20' 5" 34'
55'
(i) Testing of End B

End A

Air jack
10'
34' 20' 6"
55'
(ii) Testing of End A

(b) AASHTO BT-54


P

End B

Air jack
7' 9"
15' 6" 39'
(i) Testing of End B
P

End A
Air jack
7' 9"
39' 15'6 "
55'
(ii) Testing of End A
(c) AASHTO Type III-a, AASHTO Type III-b, and Nebraska NU-1100

P
7' 9"

31'

32'
(i) Testing of End B
P
7' 9"

31'

32'
(ii) Testing of End A
(d) Texas U-40

Figure 3.4.   Loading arrangements.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Experimental Research Approach, Findings, and Associated Analytical Simulations   61  

Figure 3.5.   Extension of cracking into End A span in


AASHTO BT-54.

3.5.2 Instrumentation
During fabrication of the girders, a number of electrical resistance strain gages were bonded
to transverse and longitudinal reinforcing bars. Moreover, a number of electrical resistance
strain gages were bonded to the second-layer strands before casting AASHTO BT-54, AASHTO
Type III-a, and AASHTO Type III-b. After the girders were delivered to the University of
Cincinnati Large Scale Test Facility, additional strain gages were bonded to a number of the
strands. These additional gages were applied within small “knockouts” left in the girders dur-
ing concrete placement in a procedure used for all specimens tested in this program. At each
end, six strain gages were bonded to the concrete surface to monitor compressive strain near the
top of the bridge deck. Five vibrating wire gages were placed in the concrete near the centroid

Figure 3.6.   Post-tensioning of cracked end before testing End A of AASHTO BT-54.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

62   Strand Debonding for Pretensioned Girders

of the strands at each end. The locations and numbers of strain gages are summarized in
Appendix G. The test specimens were externally instrumented to measure the slip of a number
of bonded and debonded strands (the locations at which the slips were measured are provided
in Appendix G), average shear deformation within the shear span, and the deflection at the
load point. The vertical displacement of the girder at the center of each support was measured
in order to account for the deformation of the neoprene pads. A calibrated pressure transducer
was used to monitor the applied load from the hydraulic rams.

3.5.3  Test Results and Discussions


With the exception of Nebraska NU-1100 and End B of Texas U-40, all the specimens were
loaded to failure. Total failure at End B of Texas U-40 would have compromised, if not effec-
tively prevented, the testing of End A. Therefore, End B of this girder was loaded to only its pre­
dicted capacity, which will be discussed in Section 3.5.3.1. It was deemed unsafe to load the
Nebraska NU-1100 girder, having 22 0.7-in. diameter strands, to failure considering the amount
of energy that would have been released in the event of a catastrophic failure. This girder, there-
fore, was loaded to only slightly above its predicted capacity.

3.5.3.1  Capacity, Stiffness, and Failure Mode


The strains measured by the vibrating wire gages were used to infer the magnitude of prestress
loss. The total losses ranged between 3% for Texas U-40 and 11% for Nebraska NU-1100. These
values, in conjunction with the measured material properties, were used to calculate the expected
capacity of each specimen per AASHTO LRFD Bridge Design Specifications (see Appendix H). In
the calculations, the resistance factors were taken as unity since the test girders were cast under
controlled conditions, the loading was well defined and known a priori, and the purpose of the
calculation was to determine a predicted capacity, not a design load. The measured loads (and
shears) were normalized with respect to the calculated capacities of each girder. The measured
deflections were normalized with respect to the deflection measured at the calculated capacity.
The resulting normalized load-deflection responses are illustrated in Figure 3.7. Table 3.7 com-
pares the normalized peak loads and normalized deflections at peak load for End A and End B.
In each case, End B met the current AASHTO limits on the amount of debonding, while End A
exceeded these limits.
Based on the presented results, the following observations are made:
1. All the specimens successfully developed their predicted capacities. The failure loads were at
least 43% larger than the nominal capacities (no reduction factors) calculated based on the
measured material properties and inferred prestress loss. The normalized failure loads were
on the order of 20% greater when determined based on AASHTO LRFD Bridge Design Speci-
fications and nominal material properties. The large amounts of debonding at End A were not
detrimental to the expected load-carrying capacity of the girders.
2. With the exception of AASHTO BI-36, the normalized deflections at peak load for End A and
End B are comparable. The normalized load-deflection for AASHTO BI-36 [Figure 3.7(a)]
clearly illustrates that End A of this girder achieved its peak capacity at a larger deflection.
3. At peak load, which corresponds to failure if the specimen were loaded to its ultimate capac-
ity, the deflection was at least 2.3 times that when the predicted capacity was developed. The
large amounts of debonding did not negatively impact the “ductility” inherent in the pre-
stressed girders.
4. The slopes of the normalized load-deflection relationships at End A and End B are essentially
the same up to developing the predicted capacities (i.e., when the value of the normalized
load is equal to 1). The larger amount of debonding at End A did not have a noticeable effect

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Experimental Research Approach, Findings, and Associated Analytical Simulations   63  

1.6 1.6

1.4 1.4
Applied load/Load at AASHTO capacity

Applied load/Load at AASHTO capacity


1.2 1.2

1.0 1.0

0.8 0.8

0.6 0.6

End A End A
0.4 0.4
End B End B
0.2 0.2

0 0
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6
Deflection under point of application of load/Deflection measured at AASHTO capacity Deflection under point of application of load/Deflection measured at AASHTO capacity

(a) AASHTO BI-36 (b) AASHTO BT-54


1.8 1.8

1.6 1.6
Applied load/Load at AASHTO capacity

Applied load/Load at AASHTO capacity


1.4 1.4

1.2 1.2

1.0 1.0

0.8 0.8

0.6 0.6
End A End A
0.4 End B 0.4 End B

0.2 0.2

0 0
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
Deflection under point of application of load/Deflection measured at AASHTO capacity Deflection under point of application of load/Deflection measured at AASHTO capacity

(c) AASHTO Type III-a (d) AASHTO Type III-b


1.4 1.6

1.4
Applied load/Load at AASHTO capacity

Applied load/Load at AASHTO capacity

1.2

1.2
1.0
1.0
0.8
0.8
0.6
0.6
0.4 End A
End A 0.4
End B
End B
0.2 0.2

0 0
0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 0 1 2 3 4 5
Deflection under point of application of load/Deflection measured at AASHTO capacity Deflection under point of application of load/Deflection measured at AASHTO capacity

(e) Nebraska NU-1100 (f) Texas U-40

Figure 3.7.   Normalized load-deflection responses.

on the overall stiffness of the girders. This observation should be expected, as the relatively
small area of prestressing reinforcement does not affect the stiffness; and debonding, which
is localized near the girder ends, has little or no effect on deflection.
The failure patterns of the girders loaded to their ultimate capacity are summarized in Fig-
ure 3.8. Based on these photographs, the failure modes were characterized as noted in this figure.
The Nebraska NU-1100 and End B of Texas U-40 were not loaded to failure. The dowel action
at End A of AASHTO Type III-b (shown in Figure 3.9) is believed to account for the residual
strength following the initial loss of carrying capacity that is apparent in Figure 3.7(d).

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

64   Strand Debonding for Pretensioned Girders

Table 3.7.    Comparison of normalized peak load


and deflection at peak load.

Normalized Deflection
Normalized Peak Load
Girder at Peak Load

End A End B End A End B


AASHTO BI-36 1.43 (1.84) 1.52 (2.04) 4.58 2.31
AASHTO BT-54 1.51 (1.95) 1.50 (1.92) 3.68 3.37
AASHTO Type III-a 1.69 (2.00) 1.78 (2.05) 5.64 5.47
AASHTO Type III-b 1.63 (1.97) 1.63 (1.92) 4.15 4.62
Nebraska NU-1100 1.21* (1.64) 0.96* (1.21) 1.37 1.00*
Texas U-40 1.53 (1.80) 1.00* (1.17) 3.03 1.01*
*Not loaded to failure.
( ) capacity ratio based on the nominal material properties, all applicable.
AASHTO reductions, and prestressing loss determined per AASHTO
LRFD Article 5.9.5.1.

3.5.3.2  Crack Patterns


Photo collages of each girder at failure, shown in Figure 3.10, were assembled to determine
the angles of diagonal cracks. These collages generally do not indicate any discernible differences
between the crack patterns at either end of a given girder. However, for some girders (e.g., Type III-a
and Type III-b), End A experienced more cracking and exhibited more of a flexure-shear behavior
than End B whose behavior was predominately controlled by web shear. These observations are
consistent with the smaller amount of prestressing force (due to greater debonding) at End A.

Failure mode: Shear compression


End A

Failure mode: Shear compression


End B
(a) AASHTO BI-36

Figure 3.8.   Failure patterns and modes of failure.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Experimental Research Approach, Findings, and Associated Analytical Simulations   65  

Failure mode: Shear tension


End A

Failure mode: “Sliding shear” at the web-flange interface


End B
(b) AASHTO BT-54

Failure mode: Shear compression


End A

Failure mode: Shear compression


End B
(c) AASHTO Type III-a

Figure 3.8.  (Continued).

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Failure mode: Shear compression


End A

Failure mode: Shear compression


End B
(d) AASHTO Type III-b

Failure mode: Shear tension and bearing


End A
(e) Texas U-40
End B was not loaded to failure

Figure 3.8.  (Continued).

Figure 3.9.   Dowel action evident in AASHTO Type III-b.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Experimental Research Approach, Findings, and Associated Analytical Simulations   67  

North face

South face
Max. dr = 0.50 Max. dr = 0.18
End A End B
(a) AASHTO BI-36

North face

South face
Max. dr = 0.60 Max. dr = 0.10
End A End B
(b) AASHTO BT-54

Figure 3.10.   Photo collages of crack patterns.


(continued on next page)

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

68   Strand Debonding for Pretensioned Girders

North face

South face
Max. dr = 0.50 Max. dr = 0.25
End A End B
(c) AASHTO Type III-a

North face

South face
Max. dr = 0.56 Max. dr = 0.22
End A End B
(d) AASHTO Type III-b

Figure 3.10.  (Continued).

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Experimental Research Approach, Findings, and Associated Analytical Simulations   69  

North face

South face
Max. dr = 0.45 Max. dr = 0.27
End A End B
(e) Nebraska NU-1100

North face

South face
Max. dr = 0.50 Max. dr = 0.23
End A End B
(f) Texas U-40

Figure 3.10.  (Continued).

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

70   Strand Debonding for Pretensioned Girders

As evident from Table 3.8, the average crack angles were essentially the same for the two ends
of a single girder having different debonding ratios. The crack widths at End A, which had a
larger debonding ratio than End B, were generally slightly wider than those at End B. However,
the maximum measured crack widths corresponding to the AASHTO-predicted capacities are
small; the largest crack width was less than 0.03 in. The larger dr did not have a deleterious effect
on observed crack angles or crack widths.

3.5.3.3  Shear Deformation


Using the displacements measured by the diagonal displacement transducers (see Appendix G),
the average shear deformations in two adjacent regions were obtained: Region 1 is approximately
one-half the shear span closer to the support, and Region 2 is the other half of the shear span closer
to the applied load. The relationship between the normalized shear and average shear strains in
these regions is shown in Figure 3.11. In this figure, the normalizing “shear at AASHTO capacity”
refers to the shear capacity determined using measured material properties, prestress loss inferred
from the strains measured by the vibrating wire gages, and taking the resistance factors as
unity. The diagonal sensors used to obtain this data were removed prior to reaching the failure
load. The load at which the displacement transducers were removed was not always identical for
End A and End B in the same girder.
In general, the shear strain for a given value of applied shear at End A was larger than that at
End B. The smaller amount of prestressing force (resulting from the larger dr) at End A could
not restrain the growth and widening of the cracks as well as in End B.

3.5.3.4  Shear Resistance from Transverse Reinforcement


The measured stress-strain relationships were used to infer stresses in the transverse rein-
forcement. The measured relationships are presented in Appendix F and are modeled in the
more appropriate of two ways as indicated in Appendix F. If the measured stress-strain relation-
ships exhibited a well-defined yield point, a trilinear stress-strain model, such as that shown in
Figure 3.12(a), was used. The values defining the model ( fy, Es, Esh, ey, and esh) were obtained
based on the data from material testing (Appendix F).
In the absence of a well-defined yield point, a Ramberg-Osgood (R-O) (Ramberg and Osgood
1943) function [Figure 3.12(b) and Eq. 3.5] was calibrated to fit the experimentally obtained
stress-strain relationships of the reinforcing steel. This continuous function is more precise than
the conventional elastic-perfectly plastic assumptions used in design.

1− A
f ss = Es ε  A + ≤ f
C 1C 
pu Eq. 3.5
 
1 + ( Bε ) 
 

Table 3.8.    Average measured angles of diagonal cracks.

End A End B
Girder
Max. dr (deg.) wmax (in.) Max. dr (deg.) wmax (in.)
AASHTO BI-36 0.50 29 0.01 0.18 30 0.01
AASHTO BT-54 0.60 34 0.025 0.10 32 0.022
AASHTO Type III-a 0.50 35 0.028 0.25 35 0.014
AASHTO Type III-b 0.56 33 0.015 0.22 34 0.025
Nebraska NU-1100 0.45 32 * 0.27 32 0.015
Texas U-40 0.50 32 0.014 0.23 34 0.01
: Average angle of diagonal cracks measured at the conclusion of testing.
wmax: Maximum crack width at a load nearly equal to the AASHTO-predicted girder capacity.
*: Not measured.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Experimental Research Approach, Findings, and Associated Analytical Simulations   71  

1.6 1.6
Applied shear/Shear at AASHTO capacity

Applied shear/Shear at AASHTO capacity


1.4 1.4

1.2 1.2

1.0 1.0

0.8 0.8

0.6 0.6

0.4 0.4
End A (Region 1) End B (Region 1) End A (Region 1) End B (Region 1)
0.2 End A (Region 2) End B (Region 2) 0.2 End A (Region 2) End B (Region 2)

0 0
0 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45 0 0.05 0.10 0.15
Average shear strain (%) Average shear strain (%)
(a) AASHTO BI-36 (b) AASHTO BT-54
1.8 1.8

1.6 1.6
Applied shear/Shear at AASHTO capacity

Applied shear/Shear at AASHTO capacity


1.4 1.4

1.2 1.2

1.0 1.0

0.8 0.8

0.6 0.6

0.4 0.4
End A (Region 1) End B (Region 1) End A (Region 1) End B (Region 1)
0.2 End A (Region 2) End B (Region 2) 0.2 End A (Region 2) End B (Region 2)

0 0
0 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0 0.05 0.10 0.15 0.20 0.25 0.30 0.35
Average shear strain (%) Average shear strain (%)
(c) AASHTO Type III-a (d) AASHTO Type III-b

1.4 1.6
Applied shear/Shear at AASHTO capacity

Applied shear/Shear at AASHTO capacity

1.2 1.4

1.2
1.0
1.0
0.8
0.8
0.6
0.6
0.4
0.4
End A (Region 1) End B (Region 1) End A (Region 1) End B (Region 1)
0.2 0.2
End A (Region 2) End B (Region 2) End A (Region 2) End B (Region 2)

0 0
0 0.05 0.10 0.15 0.20 0 0.05 0.10 0.15
Average shear strain (%) Average shear strain (%)
(e) Nebraska NU-1100 (f) Texas U-40

Figure 3.11.   Normalized applied shear vs. average shear strain.

Where:
A, B, and, C = Parameters established from a best fit of experimental stress-strain data. The
values of these parameters are summarized in Appendix F.
fss = Steel stress
fpu = Ultimate strength
Es = Modulus of elasticity usually taken as 29,000 ksi
e = Strain

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

72   Strand Debonding for Pretensioned Girders

Stress, fss Stress, fss

E(1-A)/B
AEs
fy Esh

C is a measure of
Es
transition "roundness"
Es
y sh Strain, Strain,
(a) Trilinear idealization of stress-strain (b) Ramberg-Osgood idealization of stress-
diagram strain diagram

Figure 3.12.   Modeling of stress-strain relationships for steel reinforcing.

For AASHTO BT-54 and Texas U-40, the R-O model in Figure 3.12(b) was used while the
stress-strain relationships for the transverse steel in the other girders were characterized based
on the elastic-plastic model shown in Figure 3.12(a).
The calibrated equation was used to infer stresses corresponding to the strains measured by
the strain gages bonded to the transverse reinforcement. Based on Article 5.8.3 in AASHTO
LRFD Bridge Design Specifications, the shear resistance provided by the transverse steel, Vs, was
computed. AASHTO LRFD Eq. 5.8.3.3-4 was modified slightly by using the experimentally
inferred stress ( fv) instead of the yield strength of transverse reinforcement ( fvy), as shown in
Eq. 3.6.

Av f v dv ( cot θ + cot α ) sin α Av f v dv ( cot θ )


Vs = = Eq. 3.6
s s

The value of q was selected based on the values tabulated in Table 3.8. This calculation was
performed at six locations (1, 2, 3, 4, 5, and 6 ft from the ends of the girder) where the transverse
reinforcement had been instrumented. Since there was no inclined prestressing strand, the dif-
ference between the applied shear and the average of Vs from these six locations corresponds
to the concrete contribution to shear resistance. The resulting concrete contribution to shear
capacity as a function of deflection under the load point is plotted in Figure 3.13. For a given
value of deflection, the concrete shear resistance at End A (greater dr) is less than its counterpart
in End B. This observation is consistent with the differences in the amount of prestressing force
at the two ends. The smaller prestressing force at End A resulted in more cracking and hence a
reduction in the contribution of the concrete to the shear resistance, as evident from Table 3.9.
This reduction at the AASHTO-predicted capacity is not, however, proportional to the relative
magnitude of drs at the two ends. For instance, the concrete at End A resisted 15% less shear than
End B in AASHTO BT-54, which had the greatest difference between the drs at the two ends, but
AASHTO Type III-b exhibited the largest apparent reduction in concrete contribution (19%)
even though the difference between the drs at its two ends was less substantial.

3.5.3.5  Apparent Strand Slip


Displacement transducers measured the movements of the instrumented strands relative to
the end surface of the girder (see Appendix G). For fully bonded strands, this movement is the
actual slip. In the case of debonded strands, the strand is assumed to be unstressed between the
point of measurement and the beginning of strand embedment (at 3, 6, 9, or 12 ft into the beam);
thus, the debonded portion of strand is moving as a rigid body. However, due to flexure- and
shear-induced tensile strains, the concrete mass between the end of the girder and the begin-
ning of strand embedment is elongating, i.e., ecl in Figure 3.14 where ec = concrete longitudinal

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

240 400

360
210
Shear resistance from concrete (kips)

Shear resistance from concrete (kips)


320
180
280
150
240

120 200

160
90
120
60
End A 80 End A
End B End B
30
40

0 0
0 0.5 1.0 1.5 2.0 0 0.6 1.2 1.8
Deflection under point of application of load (in.) Deflection under point of application of load (in.)
(a) AASHTO BI-36 (b) AASHTO BT-54

280 250
Shear resistance from concrete (kips)

Shear resistance from concrete (kips)


200
210

150

140

100

70
End A 50 End A
End B End B

0 0
0 0.5 1.0 1.5 2.0 2.5 0 0.5 1.0 1.5 2.0 2.5
Deflection under point of application of load (in.) Deflection under point of application of load (in.)
(c) AASHTO Type III-a (d) AASHTO Type III-b

280 400

360
240
Shear resistance from concrete (kips)

Shear resistance from concrete (kips)

320
200 280

240
160
200
120
160

80 120

80 End A
End A
40 End B
End B
40

0 0
0 0.15 0.30 0.45 0.60 0 0.5 1.0 1.5 2.0
Deflection under point of application of load (in.) Deflection under point of application of load (in.)
(e) Nebraska NU-1100 (f) Texas U-40

Figure 3.13.   Concrete shear resistance vs. deflection.

Table 3.9.    Normalized concrete shear resistance


at AASHTO-predicted girder capacities.

End A End B % Reduction in


Girder
Max. dr Vc / f'cbvdv Max. dr Vc / f'cbvdv Vc / f'cbvdv
AASHTO BI-36 0.50 0.19 0.18 0.20 5%
AASHTO BT-54 0.60 0.17 0.10 0.20 15%
AASHTO Type III-a 0.50 0.14 0.25 0.16 13%
AASHTO Type III-b 0.56 0.13 0.22 0.16 19%
Nebraska NU-1100 0.45 0.23 0.22 0.23 0%
Texas U-40 0.50 0.24 0.22 0.26 8%

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

74   Strand Debonding for Pretensioned Girders

Figure 3.14.   Conceptual illustration of elongation of


debonded strands.

strain and l = the length over which the strand is debonded. The slip measured at the end of
the girder on debonded strands will, therefore, be greater than the actual slip exhibited at the
beginning of strand embedment (a distance l into the girder). The difference between measured
slips and actual slips will, therefore, be proportional to the unbonded length (Hypothesis A).
This proportionality is unlikely to be linear since the strains concerned are not uniform over the
debonded lengths.
The strains causing the concrete deformation can only be assessed in a very general fashion
because the available data (the strains measured by the vibrating wire gages every 1 ft up to 5 ft
into the girder) did not have sufficient resolution. It was, therefore, not deemed appropriate to
attempt to correct measured slip values to account for ec l. The strains will also be affected by the
location of the strand in the section (since a strain gradient is present) and by the extent and
pattern of local cracking. Nonetheless, in a broad sense it may be hypothesized (Hypothesis B)
that the concrete strains will be greater at End A because of the smaller prestressing force in com-
parison to End B. The over-estimation of the actual slip at the beginning of strand embedment
will, therefore, be greater at End A.
The relationships between the normalized applied shear and apparent slip of bonded and
debonded strands having various debonding lengths are plotted in Figure 3.15. The values of
apparent slip at the AASHTO-predicted capacity are summarized in Table 3.10.
Based on the presented data, the following observations about “measured slip” may be drawn:
1. At AASHTO-predicted girder capacities, measured slip rarely exceeded 0.04 in., except for
Texas U-40. The measured slip of fully bonded strands was negligible in all cases.
2. In all cases, although the measured slip was negligible, the effect of specimen initial cracking
is evident as a change in slope of the normalized shear-apparent slip curves.
3. Prior to initial cracking, the slope of the slip behavior is inversely proportional to the debond-
ing length. This observation is consistent with Hypothesis A.
4. In all but AASHTO BT-54, considering End A strands, the measured slip is typically propor-
tional to the unbonded length, which is consistent with Hypothesis A. The measured slips of
AASHTO BT-54 are not entirely consistent: the strand having 9 ft debonding has the lowest
measured slip of the debonded strands.
5. Beyond the AASHTO-predicted capacity, measured slip is observed to increase as cracking
and, presumably, yield of strand and embedded steel takes place.
6. Comparing the post-AASHTO capacity of the fully bonded strands, End A is seen to exhibit
greater slip at comparable load levels, which supports Hypothesis B.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

1.6 1.6
Applied shear/Shear at AASHTO capacity

Applied shear/Shear at AASHTO capacity


1.4 1.4

1.2 1.2

1.0 1.0

0.8 0.8

0.6 0.6
End A (bonded) End A (debonded 12')
0.4 End A (debonded 3') End B (bonded) 0.4
End A (debonded 6') End B (debonded 3') End A (bonded) End A (debonded 9')
0.2 End A (debonded 9') End B (debonded 6') 0.2 End A (debonded 3') End B (bonded)
End A (debonded 6') End B (debonded 3')
0 0
0 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45 0.50 0.55 0 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45 0.50
Apparent slip (in.) Apparent slip (in.)
(a) AASHTO BI-36 (b) AASHTO BT-54

1.8 1.8

1.6 1.6
Applied shear/Shear at AASHTO capacity

Applied shear/Shear at AASHTO capacity


1.4 1.4

1.2 1.2

1.0 1.0

0.8 0.8

0.6 0.6
End A (bonded) End B (bonded) End A (bonded) End B (bonded)
0.4 0.4
End A (debonded 3') End B (debonded 3') End A (debonded 3') End B (debonded 3')
0.2 End A (debonded 6') End B (debonded 6') 0.2 End A (debonded 6') End B (debonded 6')
End A (debonded 9') End A (debonded 9')
0 0
0 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45 0.50 0.55 0.60 0 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45 0.50 0.55 0.60
Apparent slip (in.) Apparent slip (in.)
(c) AASHTO Type III-a (d) AASHTO Type III-b

1.4 1.6
Applied shear/Shear at AASHTO capacity

Applied shear/Shear at AASHTO capacity

1.2 1.4

1.2
1.0
1.0
0.8
0.8
0.6
0.6
0.4 End A (bonded) End A (debonded 12')
End A (debonded 3') End B (bonded) 0.4 End A (bonded) End B (bonded)
End A (debonded 6') End B (debonded 3') End A (debonded 3') End B (debonded 3')
0.2 End A (debonded 9') End B (debonded 12') 0.2 End A (debonded 6') End B (debonded 6')
End A (debonded 9') End B (debonded 9')
0 0
0 0.025 0.050 0.075 0.100 0 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45
Apparent slip (in.) Apparent slip (in.)
(e) Nebraska NU-1100 (f) Texas U-40

Figure 3.15.   Normalized shear-apparent slip relationships.

Table 3.10.    Apparent strand slip at AASHTO-predicted capacity.

Debonded Strands
Bonded Strands
Girder ldebond = 3 ft ldebond = 6 ft ldebond = 9 ft ldebond = 12 ft
End A End B End A End B End A End B End A End B End A End B
AASHTO BI-36 < 0.001 < 0.001 0.016 0.004 0.019 < 0.001 0.016 N/A 0.002 N/A
AASHTO BT-54 0.009 0.004 0.041 0.008 0.043 N/A 0.023 N/A N/A N/A
AASHTO Type III-a 0.004 < 0.001 0.005 0.002 0.011 0.006 0.010 N/A N/A N/A
AASHTO Type III-b < 0.001 < 0.001 0.008 < 0.001 0.018 0.006 0.022 N/A N/A N/A
Nebraska NU-1100 0.005 0.002 0.033 0.011 0.035 N/A 0.039 N/A 0.040 0.030
Texas U-40 0.011 0.003 0.036 0.060 0.073 0.034 0.103 0.099 N/A N/A

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

76   Strand Debonding for Pretensioned Girders

7. Comparing the measured slip of the End A and End B strands having unbonded lengths
of 3 ft (the only strands available for such comparison), End A, having greater debonding,
exhibits greater slip except for Texas U-40. This observation supports Hypothesis B.
Thus, both Hypotheses A and B are supported by experimental observations: (A) the difference
between measured slips and actual slips is proportional to the unbonded length; and (B) the con­
crete strains will be greater at End A because of the smaller prestressing force in comparison
to End B.

3.5.3.6  Contribution of Longitudinal Reinforcement


The amount of required nonprestressed longitudinal reinforcement at the critical section
and at the interior face of the support had been determined according to AASHTO LRFD
Eq. 5.8.3.5-1 and Eq. 5.8.3.5-2., respectively:

Mu N u  Vu 
Aps f ps + As f y ≥ + 0.5 + − Vp − 0.5Vs  cot θ [5.8.3.5–1]
dv φ f φc  φ v 

 Vu 
Aps f ps + As f y ≥ +  − 0.5Vs − Vp  cot θ [5.8.3.5–2 ]
 φv 

Based on a similar procedure discussed in Section 3.5.3.4, stresses in the nonprestressed lon-
gitudinal reinforcement were inferred from the measured strains. The R-O function given by
Eq. 3.5 and shown in Figure 3.12b was used for all the girders except for AASHTO BI-36 and
Texas U-40 girder, for which the nonprestressed longitudinal reinforcement stress-strain
relationships were based on the trilinear function depicted in Figure 3.12a. The resulting
stresses normalized with respect to the nonprestressed reinforcement yield strength are plotted
in Figure 3.16 against the normalized applied shear. If available, stresses at three sections are
plotted: (1) at the critical section near the support, (2) dv from the interior face of the support,
and (3) at the point where the load was applied. The following observations are made:
1. At AASHTO-predicted capacity (i.e., when the normalized shear is unity), the stress in the
nonprestressed reinforcing steel is at most 0.56fy, as can be seen more clearly from Table 3.11.
However, the longitudinal nonprestressed reinforcement is assumed to have yielded accord-
ing to AASHTO LRFD Eq. 5.8.3.5-1 and Eq. 5.8.3.5-2. A plausible explanation for this dif-
ference could be that AASHTO LRFD Bridge Design Specifications do not account for the
tensile strength of the precompressed concrete. Hence, the available capacity (in the absence
of nonprestressed reinforcement) is larger than Aps fps alone.
2. For the girders that were loaded to failure, the nonprestressed longitudinal bars had begun to
yield at sections 2 and 3, i.e., at the critical section near the support and dv from the interior
face of the support, respectively.

3.6 Summary
The experimentally determined girder capacities exceed those computed based on AASHTO
LRFD Bridge Design Specifications using the measured material properties and prestress losses
with no strength reduction factor. Regardless of the drs, the measured deflection at the peak
load was several times larger than the measured deflection at the calculated AASHTO capacity.
The nonprestressed reinforcement used to compensate for larger prestressed reinforcing drs
is adequate in terms of capacity. Even though this reinforcement cannot replicate the effects of
prestressing force in bonded strands, the differences in the overall stiffness, crack widths, and

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

1.6 1.6
Applied shear/Shear at AASHTO capacity

Applied shear/Shear at AASHTO capacity


1.4 1.4

1.2 1.2

1.0 1.0

0.8 End A (Section 2) 0.8 End A (Section 2) End B (Section 2)


End A (Section 3)
End A (Section 3) End B (Section 3)
0.6 End A (Section 4) 0.6 End A (Section 4)
Section 2: critical section near support Section 2: critical section near support
0.4 Section 3: dv from the face of support 0.4
Section 3: dv from the face of support
Section 4: at application of load Section 4: at application of load
0.2 0.2
End B did not have nonprestressed reinforcement Nonprestressed reinforcement at End B was terminated prior to Section 4.
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1
Stress in nonprestressed reinforcement/fy Stress in nonprestressed reinforcement/fy
(a) AASHTO BI-36 (b) AASHTO BT-54

1.8 1.8

1.6 1.6
Applied shear/Shear at AASHTO capacity

Applied shear/Shear at AASHTO capacity


1.4 1.4

1.2 1.2

1.0 1.0

0.8 0.8
End A (Section 2) End B (Section 2) End A (Section 2) End B (Section 2)
0.6 End A (Section 3) End B (Section 3) 0.6 End A (Section 3) End B (Section 3)
End A (Section 4) End B (Section 4) End A (Section 4) End B (Section 4)
0.4 0.4
Section 2: critical section near support Section 2: critical section near support
Section 3: dv from the face of support Section 3: dv from the face of support
0.2 0.2
Section 4: at application of load Section 4: at application of load
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1
Stress in nonprestressed reinforcement/fy Stress in nonprestressed reinforcement/fy
(c) AASHTO Type III-a (d) AASHTO Type III-b

1.4 1.6
Applied shear/Shear at AASHTO capacity

Applied shear/Shear at AASHTO capacity

1.2 1.4

1.2
1.0
1.0
0.8
0.8
0.6 Strain gages at Section 4 malfunctioned.
0.6
End A (Section 2) End B (Section 2)
0.4 End A (Section 2) End B (Section 2)
End A (Section 3) End B (Section 3)
End A (Section 3) End B (Section 3) 0.4
Section 2: critical section near support
0.2 Section 2: critical section near support 0.2 Section 3: dv from the face of support
Section 3: dv from the face of support Section 4: at application of load
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Stress in nonprestressed reinforcement/fy Stress in nonprestressed reinforcement/fy
(e) Nebraska NU-1100 (f) Texas U-40

Figure 3.16.   Normalized stress in nonprestressed longitudinal reinforcement.

Table 3.11.    Normalized stress in nonprestressed longitudinal


reinforcement at AASHTO-predicted girder capacities.

End A End B
Girder fs /fy fs /fy
Max. dr Max. dr
Section 2 Section 3 Section 4 Section 2 Section 3 Section 4
AASHTO BI-36 0.50 0.03 0.11 0.21 0.18 N/A
AASHTO BT-54 0.60 0.12 0.23 0.15 0.10 0.17 0.20 N/A
AASHTO Type
0.50 0.12 0.14 0.15 0.25 0.04 0.11 0.16
III-a
AASHTO Type
0.56 0.07 0.23 0.27 0.22 0.20 0.08 0.03
III-b
Nebraska NU-1100 0.45 0.56 0.10 N/A 0.27 0.31 0.09 N/A
Texas U-40 0.50 0.35 0.28 N/A 0.23 0.30 0.18 N/A

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

78   Strand Debonding for Pretensioned Girders

crack patterns and angle of cracks of the two girder ends, with different magnitudes of drs, were
found to be small.
The results are consistent with the hypothesis that bonded strand and nonprestressed tension
reinforcement work together to resist longitudinal forces induced by shear (i.e., those calculated
using AASHTO LRFD Eq. 5.8.3.5-1 and -2 provided the detailing rules shown in Table 3.2 and
elaborated upon in Section 4.2 are satisfied).

3.7  Modeling of Test Specimens


The FEM platform described in Section 2.4 and STM modeling described in Section 2.5 were
applied to the test girders in order to both validate the analytical approaches and to gain a better
understand of observed experimental behavior.

3.7.1  FEM Simulation of Test Girders


Reinforcing steel and concrete material properties were selected based on those determined
experimentally for each girder. Due to the test times, concrete properties vary from End A to
End B; this variation was captured in the models. The material properties used to model each
girder are summarized in Table 3.12. The transfer length was taken as 30db, which is consid-
ered to be more realistic of in situ behavior than 60db used in AASHTO LRFD Bridge Design
Specifications.
As evident from Figure 3.17, the load-deflection responses determined from FEM predict the
experimental curves quite well. The “saw tooth” behavior of the experimental curves reflects
the relaxation of the applied load when loading was paused to inspect the girders for cracking
and checking the data. The observed and predicted crack patterns at failure are compared in
Figure 3.18. The crack patterns based on nonlinear FEM analysis replicate those observed rea-
sonably well.

3.7.2  Utilization of Calibrated Analytical FEM Platform


As evident from the results shown in Section 3.7.1, the overall measured responses of the test
girders are quite close to those predicted by the FEM platform. Using the platform, additional
analyses were conducted to (1) further evaluate the transfer lengths and (2) examine the ramifi-
cations of rebonding a large number of previously debonded strands.

3.7.2.1  Transfer Length


The distribution of the longitudinal strain at prestress force transfer (release) was computed
by FEM analysis. As shown in Figure 3.19, the FEM results are generally close to those deter-
mined using fundamental mechanics (see Section 3.4). The results for Texas U-40 girder are dis-
cussed in Section 3.7.2.2. At a few locations, the strains from FEM are closer to the measured data
in comparison to those from basic principles. Nevertheless, a number of the measured strains
do not correspond to those based on FEM analysis. The field boundary conditions include some
restraint from the forms and are therefore different from a simple span that is used in the cal-
culations and analyses.

3.7.2.2  Further Evaluation of Longitudinal Strains at Release in Texas U-40


In the Texas U-40 girder, the concrete strains measured at release were markedly smaller than
those determined from fundamental mechanics, AASHTO, or NCHRP Report 603 methods, all

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Experimental Research Approach, Findings, and Associated Analytical Simulations   79  

Table 3.12.    FEM material properties.


BI-36 BT-54 NU-1100
f’c ksi End A End B End A End B End A End B
Girder 12.6 12.2 17.4 15.2 14.0 13.2
concrete Ec ksi 6472 6434 7592 7096 6550 6513
ft psi 760 740 870 820 810 780
0.20
Slab concrete f’c ksi N/A N/A 11.4 11.2 6.9 6.1
Prestressing fpu ksi 270
strand fpi ksi 0.75fpu = 202.5
No. 3 fy ksi 65 75
transverse fu ksi N/A 97 101
steel u 0.200 0.238
No. 4 fy ksi 68 70 79
transverse fu ksi 103 107 106
steel u 0.114 0.127 0.250
No. 5 fy ksi 70
longitudinal fu ksi N/A N/A 103
steel u 0.128
No. 6 fy ksi 69 66 69
longitudinal fu ksi 108 106 109
steel u 0.144 0.132 0.120

Type III-a Type III-b Texas U-40


f’c ksi End A End B End A End B End A End B
Girder 12.6 12.2 13.8 13.2 12.8 12.0
concrete Ec ksi 6473 6454 6542 6513 6511 6304
ft psi 760 750 806 780 750 730
0.20
Slab concrete f’c ksi 7.4 6.2 6.2 5.7 5.9 5.8
Prestressing fpu ksi 270
strand fpi ksi 0.75fpu = 202.5
No. 3 fy ksi 75
transverse fu ksi 111 N/A
steel u 0.258
No. 4 fy ksi 64 71
transverse fu ksi 100 110
steel u 0.223 0.157
No. 5 fy ksi 76 67
longitudinal fu ksi 113 106
steel u 0.232 0.093
No. 6 fy ksi 68
longitudinal fu ksi N/A 110
steel u 0.145

using a plane sections assumption. The experimental strains were measured along the centerline
of the girder at approximately the mid-depth of the bottom flange corresponding to approxi-
mately location 1 shown in Figure 3.20.
As seen in Figure 3.20, the strand layout was concentrated toward the webs of the girder. Due
to the flexibility of the open section, the webs are expected to resist most of the flexural strains/
stresses at release. Furthermore, the axial strains are developed over the transfer length and are
distributed into the concrete as a diagonal strut, rather than engaging the entire cross section
immediately. Taken together, it may be expected that the strains in the flange would be notably
reduced over a much longer length of the girder. This hypothesis was tested using the calibrated
FEM model for the test specimen (see Section 3.7.1). Figure 3.21 shows a view of the longitudinal
strains in the girder soffit confirming the hypothesis. There is a distinct shear lag effect along the
girder flange.
Figure 3.22 shows the measured and FEM-predicted strains along the girder following pre-
stress transfer. The results from basic principles are also provided. The FEM-predicted strains

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

80   Strand Debonding for Pretensioned Girders

350 350

300 300
Applied load (kips)

Applied load (kips)


250 250

200 200

150 150

100 100
Experimental Experimental
50 FEM 50 FEM

0 0
0 0.5 1.0 1.5 2.0 2.5 3.0 0 0.5 1.0 1.5 2.0 2.5 3.0
Deflection under point of application of load (in.) Deflection under point of application of load (in.)
End A End B
(a) AASHTO BI-36

800 800
700 700
600 600
Applied load (kips)

Applied load (kips)

500 500
400 400
300 300
200 Experimental 200 Experimental
FEM FEM
100 100
0 0
0 0.5 1.0 1.5 2.0 0 0.5 1.0 1.5 2.0
Deflection under point of application of load (in.) Deflection under point of application of load (in.)
End A End B
(b) AASHTO BT-54

450 450
400 400
350 350
Applied load (kips)

Applied load (kips)

300 300
250 250
200 200
150 150
100 Experimental 100 Experimental
FEM FEM
50 50
0 0
0 0.5 1.0 1.5 2.0 2.5 3.0 0 0.5 1.0 1.5 2.0 2.5 3.0
Deflection under point of application of load (in.) Deflection under point of application of load (in.)
End A End B
(c) AASHTO Type III-a

Figure 3.17.   Measured vs. FEM-computed load-deflection relationships.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Experimental Research Approach, Findings, and Associated Analytical Simulations   81  

500 500

400 400
Applied load (kips)

Applied load (kips)


300 300

200 200

Experimental Experimental
100 100
FEM FEM

0 0
0 0.5 1.0 1.5 2.0 2.5 3.0 0 0.5 1.0 1.5 2.0 2.5 3.0
Deflection under point of application of load (in.) Deflection under point of application of load (in.)
End A End B
(d) AASHTO Type III-b

500 500

400 400
Applied load (kips)

Applied load (kips)

300 300

200 200

100 Experimental 100 Experimental


FEM FEM

0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6
Deflection under point of application of load (in.) Deflection under point of application of load (in.)
End A End B
(e) Nebraska NU-1100

1000 1000

800 800
Applied load (kips)

Applied load (kips)

600 600

400 400

200 Experimental 200 Experimental


FEM FEM

0 0
0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
Deflection under point of application of load (in.) Deflection under point of application of load (in.)
End A End B
(f) Texas U-40

Figure 3.17.  (Continued).

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

82   Strand Debonding for Pretensioned Girders

Test
specimen at
failure
Peak load FEM: 318 kips, Test: 297 kips FEM: 341 kips, Test: 336 kips

All predicted
cracks

Cracks >
0.004 in.

Cracks >
0.008 in.

End A End B
(a) AASHTO BI-36

Test
specimen at
failure

Peak load FEM: 630 kips, Test: 640 kips FEM: 720 kips, Test: 724 kips
All predicted
cracks

Cracks >
0.004 in.

Cracks >
0.008 in.
End A End B
(b) AASHTO BT-54

Test
specimen at
failure

Peak load FEM: 379 kips, Test: 388 kips FEM: 421 kips, Test: 446 kips

All predicted
cracks

Cracks >
0.004 in.

Cracks >
0.008 in.

End A End B
(c) AASHTO Type III-a

Figure 3.18.   Observed and FEM-predicted crack patterns.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Experimental Research Approach, Findings, and Associated Analytical Simulations   83  

Test
specimen at
failure

Peak load FEM: 395 kips, Test: 401 kips FEM: 440 kips, Test: 478 kips

All predicted
cracks

Cracks >
0.004 in.

Cracks >
0.008 in.

End A End B
(d) AASHTO Type III-b

Test
specimen at
failure

Peak load FEM: 470 kips, Test: 468 kips FEM: 350 kips, Test: 346 kips

All predicted
cracks

Cracks >
0.004 in.

Cracks >
0.008 in.

End A End B
(e) Nebraska NU-1100

Test
specimen at
failure

Peak load FEM: 890 kips, Test: 998 kips FEM: 693 kips, Test: 710 kips

All predicted
cracks

Cracks >
0.004 in.

Cracks >
0.008 in.

End A End B
(f) Texas U-40

Figure 3.18.  (Continued).

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

84   Strand Debonding for Pretensioned Girders

(a) AASHTO BI-36

(b) AASHTO BT-54

(c) AASHTO Type III-a

Figure 3.19.   Comparison of measured and computed longitudinal strains at release.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Experimental Research Approach, Findings, and Associated Analytical Simulations   85  

(d) AASHTO Type III-b

(e) Nebraska NU-1100

Figure 3.19.  (Continued).

(a) End A (b) End B

Figure 3.20.   Cross section of U-40 test girder.

End B End A
0.00008
0.00010
0.00004
0.00006
-0.00040
-0.00038
-0.00036
-0.00034
-0.00032
-0.00030
-0.00028
-0.00026
-0.00024
-0.00022
-0.00020
-0.00018
-0.00016
-0.00014
-0.00012
-0.00010
-0.00008
-0.00006
-0.00004
-0.00002
0.00000
0.00002

Figure 3.21.   Longitudinal strains along girder soffit (reverse plan view).

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

86   Strand Debonding for Pretensioned Girders

Figure 3.22.   Measured strains vs. computed values in Texas U-40.

are shown along the girder centerline (strand location 1 shown in Figure 3.20) and near the web
at strand location 14 (Figure 3.20) at increments of 23.6 in. along the girder length. The FEM
predictions capture the markedly reduced strains observed along the girder centerline, confirm-
ing the research team’s hypothesis.

3.7.2.3  STM Simulation of Test Girders


Each test girder had a single instrumented transverse tie at each end (see Appendix G). Using
the measured strains, the stresses were inferred using a procedure similar to that described in
Section 3.5.3.4. In Figure 3.23, the inferred stress normalized with respect to fy is plotted against
the normalized applied shear.
The STM was applied to both the design and as-built experimental behavior of the single-web
test beams. The model parameters and results are given in Table 3.13. As is seen, the ties provided
in the as-built beams exceeded the requirements determined from the STM. Additionally, when
experimentally determined tie yield stress values are used, the capacity of the as-built tie details
met or exceeded the predicted tie capacity demand at the ultimate observed shear (Vexp).
Using the calculated tie force corresponding to Vexp (Table 3.13), the stress in each confin-
ing reinforcement bar was obtained by assuming a uniform stress for all confining bars within

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Experimental Research Approach, Findings, and Associated Analytical Simulations   87  

1.6 1.6
Applied shear/Shear at AASHTO capacity

Applied shear/Shear at AASHTO capacity


1.4 1.4

1.2 1.2

1.0 1.0

0.8 0.8

0.6 0.6

0.4 0.4
End A End A
0.2 End B 0.2 End B

0 0
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Stress in confinement reinforcement/fy Stress in confinement reinforcement/fy
(a) AASHTO BI-36 (b) AASHTO BT-54

1.8 1.8

1.6 1.6
Applied shear/Shear at AASHTO capacity

Applied shear/Shear at AASHTO capacity

1.4 1.4

1.2 1.2

1.0 1.0

0.8 0.8

0.6 0.6

0.4 0.4 End A


End A End B
0.2 End B 0.2

0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
Stress in confinement reinforcement/fy Stress in confinement reinforcement/fy
(c) AASHTO Type III-a (d) AASHTO Type III-b

1.4
1.6
Applied shear/Shear at AASHTO capacity

Applied shear/Shear at AASHTO capacity

1.2 1.4

1.2
1.0
1.0
0.8
0.8
0.6
0.6
0.4
0.4
End A End A
0.2 0.2
End B End B

0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Stress in confinement reinforcement/fy Stress in confinement reinforcement/fy
(e) Nebraska NU-1100 (f) Texas U-40

Figure 3.23.   Stress in confinement reinforcement.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

88   Strand Debonding for Pretensioned Girders

Table 3.13.    STM of test specimen bulb confinement.


BT-54 NU-1100 Type III-a Type III-b
dr End A End B End A End B End A End B End A End B
0.60 0.11 0.45 0.38 0.50 0.30 0.56 0.29
H (in.) 60 60 49 49 51 51 51 51
N 8 18 12 16 8 12 8 14
nf 3 7 5 6 2 2 2 3
xp (in.) 7.7 6.1 9.8 9.7 8.0 8.0 8.0 7.0
yp (in.) 2.7 3.1 2.4 2.3 3.0 3.0 3.0 3.3
hb (in.) 10.5 10.5 12.7 12.7 14.5 14.5 14.5 14.5
bb (in.) 22 22 36.8 36.8 20 20 20 20
cb (in.) 6.9 6.7 10.7 11.5 7.5 8.3 7.5 7.9
0.478 0.253 0.233 0.054 0.216 0.097 0.216 0.079

Vdes, (kips) 300 341 311 277 184 201 198 236
t = Vdesign (kips) 143 86 72 15 40 20 43 19
Ties req’d (No.
7 @ 4.5 4@9 4@8 1 2 @ 25 1 2 @ 25 1
@ in.)1

Vexp. (kips) 452 511 375 277 311 357 321 383
t = Vexp (kips) 216 129 87 15 67 35 69 30
fy of ties (ksi) 70 70 79 79 64 64 64 64
Ties req’d (No. 3@
9 @ 3.4 6 @ 5.4 4@8 1 2 @ 25 4 @ 8.3 2 @ 25
@ in.)1 12.5

Ties provided1 9@3 4@3+2@6 4@3+2@6 4@3+2@6


Tie stress Figure Figure 3.23b Figure 3.23e Figure 3.23c Figure 3.23d
Initial transverse
0.7Vdes 0.85Vdes 0.3Vdes 0.4Vdes 0.5Vdes 1.0Vdes 0.95Vdes 1.1Vdes
cracking
Maximum tie
0.84fy 0.73fy 0.37fy 0.22fy 0.45fy 0.20fy 0.63fy 0.26fy
stress
1
No. 4 hoops located over a distance H/4 + Lbearing; Lbearing = 12 in. in all cases

H/4 + Lbearing. Figure 3.24 illustrates the relationship between the resulting stress (normalized with
respect to the yield strength) and the experimentally inferred stress. The correlation between the
computed and experimental data is excellent considering the complexity of assessing behavior
of confining reinforcement.
All specimens appear to have experienced some degree of transverse cracking, and the
transverse steel, which remained elastic in all cases, controlled such cracking. The nature of
the tension tie-induced longitudinal cracking is such that it is expected to propagate from the

0.9

0.8 BT-54 (End B)


Best linear fit (R2 = 0.83)
Maximum measured tie force (1/fy)

0.7
Type IIIa (End A) BT-54 (End A)
0.6

0.5 Type IIIa (End A)


NU-1100 (End B)
0.4
NU-1100 (End A)
0.3
Type IIIb (End B)
0.2
Type IIIa (End B)
0.1

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Predicted tie force (1/fy)

Figure 3.24.   Predicted vs. inferred tie stress.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Experimental Research Approach, Findings, and Associated Analytical Simulations   89  

Figure 3.25.   Soffit of NU-1100 End A following


testing.

bearing. An example is shown in Figure 3.25 showing the 36.8 in. wide soffit of NU-1100 End A.
Each transverse line is a distance of approximately H/4 (the lines are spaced at 12 in. while
H/4 = 12.25 in.). The cracks resulting from loading clearly propagate from the bearing and
extend as far as H/2 from the face of the bearing.

3.8  Web Cracking


Web cracking has long been a concern for thin-webbed prestressed concrete girders. The
AASHTO Standard Specifications for Highway Bridges (1997) required that the factored shear
force be checked against the term Vcw . Vcw was defined as the shear force at the section required to
create a maximum principal tensile stress of 0.125√f c′ in ksi units in the web at the neutral axis of
the section resisting external loads (i.e., the composite section neutral axis for composite beams,
and the non-composite neutral axis for non-composite beams). If the neutral axis fell within
the top flange, the stress was calculated at the web/top flange interface. In lieu of calculating the
principal stress, a simplified equation could be used. It was known that limiting the total shear
force to 0.125√f c′ (ksi), which was the assumed stress required to crack the web, was conservative,
but it was considered the best approach available at the time.
The first edition of the AASHTO LRFD Bridge Design Specifications did not include Vcw or
the companion Vci flexural shear strength checks. The sectional method (a.k.a. the modified
compression field theory) in the AASHTO LRFD Bridge Design Specifications was adequate to
address web-cracking issues. However, the original version of the sectional method was difficult
to use and to automate in a computer code, as it required the use of tables and iteration. In 2007,
AASHTO restored Vcw and Vci checks to the LRFD Bridge Design Specifications as the simplified
method (Article 5.8.3.4.3). The only change from the form used in the Standard Specifications
was the crack angle, q, had to be calculated when Vcw was less than Vci (the crack angle was assumed
as 45° for Vci and was assumed as 45° for Vcw in the Standard Specifications).
Concerns over web cracking remain. Web cracking is addressed in two articles of the cur-
rent AASHTO LRFD Articles 5.8.3.4.3 and 5.8.5. Article 5.8.3.4.3 is the simplified method of
determining shear resistance of concrete sections. Vcw controls near the end of the girder where
debonding, if present, occurs. Hence, it is necessary to determine whether debonding has any
influence on this calculation. In June 2016, the AASHTO Subcommittee on Bridges and Struc-
tures (SCOBS) approved a reorganization of Section 5 of the AASHTO LRFD Bridge Design Spec-
ifications for publication in 2017. As part of this reorganization, the simplified method utilizing

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

90   Strand Debonding for Pretensioned Girders

Vcw and Vci was eliminated. Consequently, the need to determine whether debonding has any
influence on the calculation of Vcw will no longer exist. However, this effect was investigated as
part of this project before it was known that the article would be deleted.
AASHTO LRFD Article 5.8.5 requires that the principal tensile stress in the web of segmen-
tal box girders be investigated. The principal tensile stress may not exceed 0.11√ f c′ ksi under
Service III loading. As part of the reorganization of Section 5, this article was extended to
apply to all prestressed concrete sections with compressive strengths used for design greater
than 10 ksi. In anticipation of this implementation in 2017, the effect of debonding on web
principal tensile stresses was investigated.
AASHTO LRFD Articles 5.8.3.4.3 and 5.8.5 were examined using the experimental data
obtained as part of this project and previous studies reported in Shahawy et al. (1993 and 1996).

3.8.1  Calculation of Principal Tensile Stress


The state of stress in the web is shown in Figure 3.26.
The normal stress, fpc, is calculated from Eq. 3.7.

Ppe Ppe e M ( y − ybnc ) M L( y − ybc )


f pc = − ( y − ybnc ) + dnc + Eq. 3.7
Anc I nc I nc Ic

In Eq. 3.7, compressive stresses are positive and tensile stresses are negative. The sign of each
term provides the correct sense of the stress.
The shear stress, v, is calculated from Eq. 3.8. (Also see Figure 3.27.)

Vdnc Qnc VLQc


v= + Eq. 3.8
t w I nc tw Ic

The principal tensile web stress is then calculated from Eq. 3.9.

2
f pc  f pc 
− ft = −   + v2 Eq. 3.9
2  2 

Note that the principal stress is shown as negative, indicating tension.

3.8.2  AASHTO LRFD Specifications Article 5.8.5


The following procedure was used to check the web stress in the test girders:
1. The stress was checked at the critical section, dv, from the face of the bearing pad. For simplic-
ity, dv was taken as 0.9de where de is the effective depth of the beam.
2. Ppe was calculated at the critical section. The area of the prestressing steel was adjusted to
v account for debonded strands. Prestress losses were estimated using strain measurements
v v
from the girders. In general, this loss was between 3% and 10%. If a debonded strand was
fpc fpc bonded before the critical section, the assumed stress in the strand was linearly interpolated
if the bonded length was less than the transfer length of 60db.
v 3. Vdnc and Mdnc were calculated assuming a concrete unit weight of 0.150 kcf. The dead load
Figure 3.26.   State of shears and moments were small compared to the applied load; hence, variation of this
web stresses. assumption would not introduce appreciable error.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Experimental Research Approach, Findings, and Associated Analytical Simulations   91  

Figure 3.27.   Definition of Q for the area


above y for the section shown.

4. VL and ML were calculated from the applied load at the time the first diagonal crack was visu-
ally observed in the experiment.
5. Composite section properties were calculated by transforming the deck into an equivalent
width of beam concrete using the modulus of elasticity computed based on concrete strength
of both the beam and slab concrete measured at the time of testing; otherwise, gross section
properties were used.
6. Using Eqs. 3.7, 3.8, and 3.9, the principal tensile stress was calculated at 3 points: (1) the inter­
face of the bottom flange and the web, (2) the interface of the top flange and the web, and
(3) the composite section neutral axis (for the non-composite BI-36, the non-composite
neutral axis was used). The maximum principal tensile stress was compared to the allowable
stress of 0.11√f c′ where f c′ is the measured beam concrete strength at the time of testing in ksi.
Table 3.14 shows the maximum principal tensile stresses at the occurrence of web crack-
ing, calculated at the three locations described in step 6. The stress at the bottom flange/web
junction did not control for any of the cases shown. When the maximum principal tensile web
stress is compared to the allowable value of 0.11√f c′, the ratio exceeds 1 for all but the BI-36 box
girder. That is, the calculated principal tensile stress at cracking in the web was greater than the
minimum allowable stress for all but the box girder. This trend should be expected. The data

Table 3.14.    Maximum principal tensile stresses for test girders (fc′ > 10 ksi).

Maximum Principal Tensile Stresses


Allowable
Max. f’c @ @ Stress Max Stress NA Stress
Girder End Neutral Web/Top @ Web/Bottom
dr (ksi) 0.11 f’c Allowable Allowable
Axis Flange Flange (ksi) (ksi)
(ksi) (ksi)
AASHTO BI-36 A 0.50 12.4 0.323 0.265 0.245 0.387 0.832 0.832
AASHTO BI-36 B 0.18 12.4 0.246 0.231 0.158 0.387 0.634 0.634
AASHTO BT- 54 A 0.60 15.0 0.564 0.616 0.304 0.428 1.44 1.32
AASHTO BT- 54 B 0.10 17.4 0.618 0.804 0.249 0.459 1.75 1.35
AASHTO Type III-a A 0.50 12.4 0.472 0.367 0.460 0.388 1.22 1.22
AASHTO Type III-a B 0.25 12.4 0.407 0.423 0.220 0.388 1.09 1.05
AASHTO Type III-b A 0.56 13.5 0.414 0.362 0.379 0.404 1.02 1.02
AASHTO Type III-b B 0.22 12.5 0.491 0.440 0.440 0.404 1.21 1.21
Nebraska NU-1100 A 0.45 13.6 0.484 0.551 0.399 0.406 1.36 1.19
Nebraska NU-1100 B 0.27 13.6 0.518 0.616 0.406 0.406 1.52 1.28
Texas U-40 A 0.50 12.0 0.506 0.473 0.303 0.381 1.33 1.33
Texas U 40 B 0.23 12.0 0.583 0.586 0.220 0.381 1.54 1.53
Controlling stress is in bold.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

92   Strand Debonding for Pretensioned Girders

for the BI-36 appear to be an anomaly and will be discussed later. Assuming the anomaly of
the BI-36 can be explained, the data would indicate that the proposed changes to Article 5.8.5
requiring the principal tensile stress be checked in the webs of all prestressed girders with design
strengths above 10 ksi would remain appropriate even for heavily debonded girders such as those
described here. Table 3.14 also shows that checking the stress only at the neutral axis is sufficient
for all but the box girders.
All of the girders in this experimental program had measured concrete strengths exceed-
ing 10 ksi. Data reported in Shahawy et al. (1993 and 1996) were used to examine the proposed
changes to the AASHTO LRFD Specifications Article 5.8.5 for cases with concrete strengths less
than 10 ksi. Using information found in the references, the principal tensile stress check was
conducted for the girders with debonded strands where cracking data was available. Based on the
previous observations, the maximum principal tensile stress was computed only at the neutral
axis of composite section. The cracking load was taken from diagrams provided in the Shahawy
et al. 1993 reference. The only concrete strength reported was an average strength of 7 ksi (Shahawy
et al. 1996). This value was used for both the girders and the slabs. Table 3.15 shows the results
of this analysis. All of the girders reported by Shahawy et al. meet the provision of the calculated
principal stress at first cracking exceeding 0.11√f c′. The smallest ratio is 1.48 and several exceed 2.0.
Hence, the proposed provision of Article 5.8.5, which does not require a principal tensile stress
check for pretensioned girders with concrete strengths below 10 ksi, appears to be acceptable.

3.8.3  AASHTO LRFD Specifications Article 5.8.3.4.3


The proposed revision to Article 5.8.5 will require a stress check all along the height of the web
since the maximum principal tensile stress may not occur at the neutral axis. The data shown in
Table 3.14, except that for the BI-36 girder, suggest that checking the stress at the neutral axis is
sufficient to prevent web cracking for girders with debonded strands.
The value of the stirrup contribution, Vs, to shear capacity depends on the angle of the crack
intersecting the stirrups. Based on a Mohr’s Circle analysis, the angle of the crack, q, can be
calculated as (Eq. 3.10):

 v 
θ = 0.5arctan  Eq. 3.10
 f pc 2 

Table 3.15.    Maximum principal tensile


stresses for Shahawy et al. girders
(fc′ < 10 ksi).
Ratio of Calculated Principal
Max. Tensile Stress at NA of
Specimen I.D.
dr Composite Section to
0.11 f’c
A2-25-3R N 0.25 2.11
A2-25-3R S 0.25 1.85
A2-50-3R N 0.50 1.78
A2-50-3R S 0.50 1.73
C0-50-R N 0.50 2.42
C0-50-R S 0.50 1.48
C1-25-R N 0.25 1.59
C1-25-R S 0.25 1.52
C1-50-R N 0.50 1.72
C1-50-R S 0.50 1.48

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Experimental Research Approach, Findings, and Associated Analytical Simulations   93  

In the current version of the AASHTO LRFD Specifications, the crack angle used for deter-
mining Vs in conjunction with Vcw is calculated from Eq. 5.8.3.4.3-4 (Eq. 3.11).

 f pc 
cot θ = 1.0 + 3  ≤ 1.8 Eq. 3.11
 fc′ 

In Table 3.16, the angle calculated from Mohr’s Circle (Eq. 3.10) and that from AASHTO
LRFD Eq. 5.8.3.4.3-4 (Eq. 3.11) are compared to the average measured crack angle. All calcu-
lations were performed at the neutral axis as the AASHTO equation was developed for crack
angles at the neutral axis, and the experimental crack angles were measured at the neutral axis.
The angle predicted by the LRFD equations agrees quite well with the angle predicted by Mohr’s
Circle, and both agree reasonably well with the measured angles. One important observation is
that the measured angles on the “A” ends do not agree as well with the predictions as the “B”
ends; the angles on the “A” ends tend to be shallower than predicted. The “A” ends all have the
higher drs.
In both the Mohr’s Circle and the AASHTO equations, increasing fpc results in a shallower
angle. A possible conclusion from the experimental data is that the crack angle on End A is shal-
lower due to a higher fpc, and the higher fpc may be caused by the prestressing force being higher
than calculated. The higher stress is likely caused by the fact that the calculation assumes that
stress in debonded strands after they are rebonded is linear over the transfer length of 60db. In
reality, the stress is not linear and the transfer length is likely less than 60db. Thus, the prestress-
ing force at the section is likely higher than assumed; this difference would be more pronounced
at End A having large drs.
When using the simplified method for design, the maximum factored shear force must be
less than Vcw in the areas where Vcw controls. Thus, Vcw + Vs should be compared to the total
shear force at failure. However, Vcw is defined as the shear force that causes the principal tensile
stress to exceed 0.125√f c′ (ksi), which is assumed to crack the web. Thus, it may be appropri-
ate also to compare Vcw to the load that causes cracking in the web. Table 3.17 presents both
comparisons. Note that the data from Shahawy et al. (1993, 1996) are not included in this
table because Vcw had been checked using the Standard Specifications that tend to produce
results that are more conservative than those obtained based on AASHTO LRFD Bridge Design
Specifications.

Table 3.16.    Crack angles for test girders.


Measured Mohr's AASHTO
Max. Measured Measured
Girder End Angle Circle (deg.) LRFD (deg.)
dr Mohr's Circle AASHTO
(deg.) Eq. 3.10 Eq. 3.11
AASHTO BI-36 A 0.50 28.8 31.3 34.2 0.92 0.84
AASHTO BI-36 B 0.18 29.7 24.8 29.5 1.20 1.01
AASHTO BT-54 A 0.60 33.9 37.5 37.9 0.90 0.89
AASHTO BT-54 B 0.10 31.5 34.3 32.9 0.92 0.96
AASHTO Type III-a A 0.50 34.7 34.7 35.9 1.00 0.97
AASHTO Type III-a B 0.25 34.9 34.3 34.0 1.02 1.03
AASHTO Type III-b A 0.56 32.6 35.5 36.0 0.92 0.90
AASHTO Type III-b B 0.22 33.6 31.3 32.8 1.07 1.02
Nebraska NU-1100 A 0.46 32.3 32.9 32.9 0.98 0.98
Nebraska NU-1100 B 0.27 31.8 31.9 31.0 1.00 1.03
Texas U-40 A 0.50 32.0 38.4 38.5 0.83 0.83
Texas U-40 B 0.23 34.0 37.4 36.7 0.91 0.94

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

94   Strand Debonding for Pretensioned Girders

Table 3.17.    Comparison of Vcw to cracking shear and measured


maximum shear.
Measured V at Calculated Vcw Measured V at Calculated
Girder End
Cracking (kips) (kips) Failure (kips) Vcw+Vs (kips)
AASHTO BI-36 A 106 106 246 205
AASHTO BI-36 B 106 135 278 255
AASHTO BT-54 A 209 127 452 241
AASHTO BT-54 B 256 154 511 291
AASHTO Type III-a A 155 119 311 215
AASHTO Type III-a B 187 132 357 235
AASHTO Type III-b A 171 123 321 219
AASHTO Type III-b B 171 145 383 253
Nebraska NU-1100 A 176 120 375* 316
Nebraska NU-1100 B 196 116 277* 301
Texas U-40 A 286 167 748 433
Texas U-40 B 345 187 532* 474
*Not loaded to failure

The data shown in Table 3.17 indicate that, for all but the BI-36 girder, the total applied shear
force at cracking exceeds Vcw, and the total shear force at failure exceeds Vcw + Vs in all cases except
End B in NU-1100; however, this girder end was not loaded to failure.
The results shown in Tables 3.16 and 3.17 indicate that no change would be needed to Arti-
cle 5.8.3.4.3. Vcw is a conservative prediction of the shear force at failure for all girders tested to
failure, and is a conservative prediction of web cracking for all but the box girders (which are
addressed below).

3.8.4  Evaluation of Data for AASHTO BI-36 Test Girder


As indicated in the two previous sections, the BI-36 box girder did not satisfy the criterion of
the maximum principal tensile stress in the web at the time of first cracking, i.e., the measured
principal stress at cracking was less than 0.11√f c′. Two plausible explanations are provided in
the following:
The box girder was the only non-composite sections tested. If Eq. 3.9 is rearranged, Eq. 12 is
obtained:

f pc ft + ( f t )
2
v= Eq. 3.12

Increasing fpc increases v, which increases the shear force (and the load) that causes cracking.
In a non-composite girder, fpc = Ppe /Anc. The other terms are bending terms, which are zero at
the neutral axis of non-composite girders. If the value of Ppe is overestimated, the cracking load
will be overestimated, or restated, for a given load; overestimating Ppe will underestimate the
principal tensile stress in the web. Increasing fpc would also increase Vcw; thus, underestimating
the loss of prestressing force would increase the value of Vcw.
Using the data from the box girder, a “what if ” analysis was done. In order to obtain the observed
cracking load, the prestress losses would have to be between 35% and 50%. This loss is clearly unre-
alistic as prestressing force losses are usually 15%–25%, and those values occur after a very long
time. The girder was 97 days old when End B was tested, and 105 days old at the time of testing of
End A; hence, the losses would be expected to be less than 15%–25%.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Experimental Research Approach, Findings, and Associated Analytical Simulations   95  

Table 3.18.    Comparison of stresses at the onset of web cracking.

AASHTO BI- AASHTO


Texas U-40
36 Type III-b
End A End B End A End B End A End B
Applied load (kips) 120 120 200 200 360 440
Peak principal stress (psi) from FEM 687 607 392 435 514 435

Near Approximately Near top


Location of peak stress from FEM N.A. N.A. N.A.
N.A. @ N.A. flange

Stress (psi) from basic principles 323 246 472 407 506 586
Location of peak stress from basic Top
N.A. N.A. N.A. N.A. N.A.
principles flange/web
Stress from basic principles/FEM
0.47 0.41 1.20 0.94 0.98 1.35
stress
N.A. = Neutral axis.

A second explanation is that the critical section is in the hollow section of the box, which is
4 in. from the solid end diaphragm. The transition from the solid end diaphragm to the hollow
section creates a disturbed region (D region) and the equations for shear and bending stresses do
not apply. This hypothesis was examined by comparing the stresses calculated from fundamental
mechanics (Mohr’s Circle) to those predicted by the calibrated FEM models. The stresses were
computed for three girders: (1) AASHTO BI-36 with end diaphragms, (2) AASHTO Type III-b
with a single web, and (3) Texas U-40 that did not have end diaphragms. The results shown in
Table 3.18 indicate that the locations of peak stresses from basic principles and FEM analysis
are essentially identical, and that basic principles are sufficiently accurate to estimate the peak
stresses for the AASHTO Type III-b and Texas U-40 but not the AASHTO BI-36. Basic prin-
ciples are based on the Bernoulli beam assumption (i.e., plane sections remain plane), which are
not applicable to D regions. For the AASHTO BI-36, the FEM model shows almost twice the
stress calculated from basic principles (Mohr’s Circle).
In areas within h of concrete end diaphragms, a more exact analysis of the web stresses is
needed to obtain appropriate values. Because this is impractical in a design situation, the calcu-
lated tensile stress should be limited to 0.08 fc′ ksi under the Service I limit state.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Chapter 4

Conclusions and Suggestions

4.1 Conclusions
Through a study combining extensive parametric, analytic, numeric, and experimental com-
ponents, a better understanding of the use of debonded prestressing strands resulted. This study
was premised on the concept that debonding prestressing strands are necessary to control stresses
resulting from prestressing force released in prestressed concrete girders. This is particularly true
with heavily prestressed girders and the adoption of larger strand diameters, where harping
strands is not always a practical or sufficient option.
The primary conclusion of this study is that debonding strands—in itself—is not detrimental
to prestressed girder performance provided the requirements for longitudinal reinforcement
(AASHTO LRFD Article 5.8.3.5—described in Section 1.3.1 of this report) to resist the additional
tension due to shear are met. This finding confirms and consolidates those of previous research;
indeed, the analyses conducted in this study have shed light on and explained some inconsistent
findings of previous work.
This study also addressed design details associated with debonded strands and girder end
regions for which a number of recommendations resulted. These recommendations are reported
in the following sections.

4.2 Suggested Detailing Guidelines for Prestressed


Girders Having Debonded Strands
Based on the reported experimental and analytical studies, the following design guidelines are
proposed. Some of the guidelines are the same as currently used when the number of strands is
not greater than 25 percent. In this section, all references are to 2016 AASHTO LRFD section
numbering.
In all girders, regardless of section shape:
• The total number of debonded strands shall not exceed 60% of the total number of strands
unless test results or successful past practices indicate that a larger percentage of strands may
be debonded.
• The number of debonded strands in any horizontal row within the bottom flange height
other than the bottom row shall not exceed 80% of the number of strands in that row.
• Tensile force in prestressing reinforcement (A ps fps) shall exceed the tensile force of the
nonprestressed reinforcement (As fs) at all sections. Development of straight and bent-up
strands as well as overhangs, if present, should be taken into account for determining the
value of fps.

96

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Conclusions and Suggestions   97  

• No more than 40% of debonded strands, or four strands, whichever is greater, shall have
the debonding terminated at any section. (No change from current AASHTO LRFD Arti­
cle 5.11.4.3).
• Satisfy AASHTO LRFD Articles 5.10.10.1, 5.10.10.2, and 5.8.3.5.

In single-web flanged sections (I-beams and bulb-tees) as summarized in Figure 4.1:


• No more than 50% of the bottom row of strands shall be debonded.
• The outermost strands in all rows located within the full-width section of the flange shall
remain bonded. Full width is understood to mean the full width of the bottom flange less a
distance accounting for the chamfer—typically 2 in. on both sides.
• With the exception of the outermost strands, strands further from the section vertical center-
line shall be debonded prior to those nearer the centerline.
• Strands in the flange within the web width should remain bonded.
• Debonded strands shall be symmetrically distributed about the vertical centerline of the cross
section of the member.
• Full flange width bearing shall be provided at supports. Full flange width bearing is not neces-
sary if a steel sole plate is provided. The width of the sole plate shall be at least one-half of
the width of the bulb.
In multi-web sections having a “bottom flange” (voided slabs, box beams, and U-beams) as
summarized in Figure 4.2:
• No more than 50% of the bottom row of strands shall be debonded.
• Bearings placed below webs not connected by an end diaphragm shall engage a width equal
to twice the extension of all webs at supports.
• For girders supported at their corners, the strands located within a width equal to twice the
extension of all the webs shall remain bonded.
• For girders supported at their corners, strands shall be debonded from the centerline outward.

Figure 4.1.    Proposed details for single-web sections.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

98   Strand Debonding for Pretensioned Girders

(a) Supported at the corners

(b) Supported across full width

Figure 4.2.    Proposed details for multi-web sections without end diaphragms.

• For girders supported across their width, debonded strands shall be uniformly distributed
across the flange width between webs.
In solid slabs:
• No more than 50% of the bottom row shall be debonded.
• For slabs uniformly supported across their width, debonded strands shall be uniformly dis-
tributed across the width.
• For slab beams supported at their corners, strands above the bearings shall not be debonded.

Proposed specification changes in the format of a Working Agenda Item based on the eighth edi-
tion of the AASHTO LRFD Bridge Design Specifications (forthcoming) are provided in Appendix I.

4.3 Suggestions Regarding Transverse Tension Ties


at STRENGTH I Ultimate Limit State
The development of tension oriented transversely across the bulb of single-web sections was
identified as a potential failure mode requiring tie reinforcement through the bulb to control
associated longitudinal cracking at the STRENGTH I limit state. The nature of the resulting
failure, however, is related to excessive transverse deformation and cracking of the bulb and
is not likely to be of catastrophic nature since reinforcement satisfying AASHTO LRFD Arti-
cles 5.10.10.1 and 5.10.10.2 contribute to the tie capacity, and in many instances will be suf-
ficient to fully resist the tie force. Furthermore, although the tie force is not a function of strand
debonding per se, the guidance for debonding patterns presented above were developed partially

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Conclusions and Suggestions   99  

Figure 4.3.    STM for confinement.

with the intent of minimizing the tie force. The following requirements were developed to design
reinforcing to fully resist the tie force:
• For single-web flanged sections, the STM shown in Figure 4.3 shall be used to determine the
required amount of tie reinforcement required, where t is the tie force to be resisted (Eqs. 4.1
and 4.2):

t = (Vu φ )(nf Nw)[ x p ( hb − yp ) + ( x p − cb ) yp ] Eq. 4.1

where

cb = (bb 2 )(1 − n f Nw ) Eq. 4.2

and dimensions are defined in Figure 4.3.


• The reinforcement determined in these requirements shall be uniformly distributed over the
length of the bearing plus a distance equal to one-quarter of the overall height of the girder
(including the composite slab if provided) toward the midspan of the girder.
In some instances, these requirements may result in impractical confinement reinforcing
details. An embedded sole plate would likely be more practical in such cases.
• The steel sole plates shall be embedded at the girder ends.
• AASHTO LRFD Articles 5.10.10.1 and 5.10.10.2 shall still be applicable when steel sole plates
are used.
Proposed specification changes in the format of a Working Agenda Item based on the eighth
edition of the AASHTO LRFD Bridge Design Specifications are provided in Appendix I.

4.4  Web Shear Cracking


Analysis of experimental results shows that using a principal tensile stress of 0.11 fc′ ksi is a
reasonable lower bound for predicting shear cracking in webs for girders with debonded strands.
However, in sections with concrete end diaphragms, the Mohr’s Circle approach of AASHTO
LRFD Article 5.9.2.3.3 does not correctly predict stresses in areas near the diaphragm, as these

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

100   Strand Debonding for Pretensioned Girders

are “disturbed” regions where the assumptions implicit in a Mohr’s Circle analysis do not
apply. Mohr’s Circle under predicts the actual web stresses in these areas. A finite element
approach was needed to correctly assess the stresses in these areas. Since a finite element is
impractical in a design situation, a reduced stress limit of 0.08 fc′ is a practical solution until
more research can provide another approach.

4.5  Suggestions for Future Research


The following is a list of recommended future research activities.
1. The effects of skewed girder ends should be investigated experimentally. The effects of both
large flange prestressing forces and the ability for debonding to mitigate these should be
considered.
2. The STM should be evaluated further through a testing program focused specifically on
transverse cracking. The stiffness of bearing and embedded sole plates should be the pri-
mary test parameters.
3. The detailing guidelines and proposed code revisions need to be examined for concrete
strengths less than 10 ksi.
4. Other girder shapes, such as voided and solid slabs, should be tested.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

References

AASHTO. (2010) LRFD Bridge Construction Specifications, American Association of State Highway and Trans-
portation Officials, Washington, D.C.
AASHTO. (2016) LRFD Bridge Design Specifications, Seventh Edition, with 2015 and 2016 Interim Revisions.
American Association of State Highway and Transportation Officials, Washington, D.C.
AASHTO. (2015) LRFD Bridge Design Specifications, Interim Revision. American Association of State Highway
and Transportation Officials, Washington, D.C.
AASHTO. (2014) LRFD Bridge Design Specifications, Seventh Edition. American Association of State Highway
and Transportation Officials, Washington, D.C.
AASHTO. (2012) LRFD Bridge Design Specifications, Sixth Edition. American Association of State Highway and
Transportation Officials, Washington, D.C.
AASHTO. (2010) LRFD Bridge Design Specifications, Fifth Edition. American Association of State Highway and
Transportation Officials, Washington, D.C.
AASHTO. (2007) LRFD Bridge Design Specifications, Fourth Edition. American Association of State Highway
and Transportation Officials, Washington, D.C.
AASHTO. (1997) Standard Specifications for Highway Bridges, Sixteenth Edition. American Association of
State Highway and Transportation Officials, Washington, D.C.
AASHTO. (1996) Standard Specifications for Highway Bridges, Interim Revisions to the Sixteenth Edition.
American Association of State Highway and Transportation Officials, Washington, D.C.
AASHTO. (1992) Standard Specifications for Highway Bridges, Fifteenth Edition. American Association of State
Highway and Transportation Officials, Washington, D.C.
AASHTO. (1989) Standard Specifications for Highway Bridges, Fourteenth Edition. American Association of
State Highway and Transportation Officials, Washington, D.C.
Abdalla, O. A., Ramirez, J. A., and Lee, R. H. (1993) Strand Debonding in Pretensioned Beams: Precast Prestressed
Concrete Bridges with Debonded Strands. Simply Supported Tests [PART 2], FHWA/INDOT/JHRP-92/25,
FHWA and Indiana Dept. of Transportation, 228 pp., Purdue, IN.
Barnes, R., Burns, N. and Kreger, M. (1999) Development Length of 0.6-Inch Prestressing Strand in Standard
I-Shaped Pretensioned Concrete Beams. Report No. FHWA/TX-02/1388-1, Center for Transportation Research,
The University of Texas at Austin, 338 pp., Austin, TX.
Baxi, A. N. (2005) Analytical Modeling of Fully Bonded and Debonded Pre-Tensioned Prestressed Concrete Members.
Ph.D. Dissertation, University of Texas.
Bentz, E. C. (2000) Response 2000. http://www.ecf.utoronto.ca/~bentz/home.shtml, accessed April 29, 2016.
Bridge Office Policies and Procedures (BOPP). (2011) Nebraska Department of Roads, Bridge Division, Lincoln,
NE, 450 pp.
Briere, V., Harries, K. A., Kasan, J., and Hager, C. (2013) Dilation Behavior of Seven-Wire Prestressing Strand—
The Hoyer Effect. Journal of Construction and Building Materials, Vol. 40, pp. 650–658.
Burgueño, R., and Sun, Y. (2011) Effects of Debonded Strands on the Production and Performance of Prestressed
Concrete Beams. Research Report to MDoT SPR NO.87346, Michigan State University.
CEB-FIP Model Code. (1990) Model Code for Concrete Structures. Thomas Telford Services Ltd., London, Great
Britain; Also published by Comité euro-international du béton (CEB), Bulletin d’Information No. 213
and 214, Lausanne, Switzerland.
Collins, M. P., and Mitchell, D. (1997) Prestressed Concrete Structures. Prentice-Hall.
Csagoly, P. (1991) A Shear Moment Model for Prestressed Concrete Beams. Florida Department of Transportation,
Tallahassee, FL.
Ghosh, S. K., and Fintel, M. (1986) Development Length of Prestressing Strands, Including Debonded Strands,
and Allowable Concrete Stresses in Pretensioned Members. PCI Journal, Sept/October 1986, pp. 38–57.

101  

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

102   Strand Debonding for Pretensioned Girders

Hamilton, H. R. (2009) Shear Performance of Existing Prestressed Concrete Bridge Girders. Final Report, University
of Florida, 160 pp.
Hanna, K. E., Morcous, G., and Tadros, M. K. (2010) Design Aids of NU I-Girder Bridges, Nebraska Department of
Roads. Project No. P322, Lincoln, NE, 115 pp.
Hawkins, N. M., and Kuchma, D. A. (2007) NCHRP Report 579: Application of LRFD Bridge Design Specifica-
tions to High-Strength Structural Concrete: Shear Provisions. Transportation Research Board of the National
Academies, Washington, D.C., 195 pp.
Jaber, F. (2016) Personal Communication, April 12, 2016.
Kaar, P. H., and Magura, D. D. (1965) Effect of Strand Blanketing on Performance of Pretensioned Girders. Journal
of the Prestressed Concrete Institute, 10(6), pp. 20–34.
Kannel, J., French, C., and Stolarski, H. (1997) Release Methodology of Strands to Reduce End Cracking in Pre-
tensioned Concrete Girders. PCI Journal, 42(1), 42–54.
Krishnamurthy, D. (1971) The Shear Strength of I-Beams with Debonded Tendons. Materiaux et Constructions,
4(4), pp. 213–218.
Larson, J., Heyen, W., and Halsey, L. (2009) Elastic Modulus Testing for SCC in Concrete Piling. Nebraska Depart-
ment of Roads, Lincoln, NE, 2 pp.
Ma, Z., Tadros, M. K., and Baishya, M. (2000) Shear Behavior of Pretensioned High-Strength Concrete Bridge
I-Girders. ACI Structural Journal, 97(1), pp. 185–192.
Moore, A. M. (2010) Shear Behavior of Prestressed Concrete U-Beams, MS Thesis, The University of Texas at
Austin, 182 pp.
Morcous, G., Hanna, K., and Tadros, M. K. (2010) Bottom Flange Reinforcement in NU I-Girders, Report. Project
No. P331, University of Nebraska, 60 pp.
Nagle, T. J., and Kuchma D. A. (2007) Nontraditional Limitations on the Shear Capacity of Prestressed Concrete
Girders. Report No. NSEL-003, Urbana-Champaign, Illinois, 184 pp.
Oliva, M. A., and Okumus, P. (2010) Finite Element Analysis of Wide flange Prestressed Girders to Understand and
Control End Cracking. Research Update 12-16-2010, University of Wisconsin, Madison.
PCI. (2011) Bridge Design Manual. Precast/Prestressed Concrete Institute, Chicago, IL.
Rabbat, B. G., Kaar, P. H., Russell, H. G., and Bruce, R. N., Jr. (1979) Fatigue Tests of Pretensioned Girders with
Blanketed and Draped Strands. PCI Journal, 24(4), July–August, pp. 88–115; Also, reprinted as PCA Research
and Development Bulletin RD062.01D, Portland Cement Association, Skokie, IL.
Ramberg, W., and Osgood, W. R. (1943) Description of Stress-Strain Curves by Three Parameters. National Advi-
sory Committee on Aeronautics, TN 902.
Ramirez, J. A., and Russell, B. W. (2008) NCHRP Report 603: Transfer, Development, and Splice Length for Strand/
Reinforcement in High-Strength Concrete. Transportation Research Board of the National Academies, Wash-
ington, D.C., 121 pp.
Ross, B. E., Hamilton, H. R., and Consolazio, G. R. (2013) Design Model for Confinement Reinforcement in Pre-
tensioned Concrete I-Girders, Transportation Research Record: Journal of the Transportation Research Board,
No. 2331, Washington, D.C., pp. 59–67. http://dx.doi.org/10.3141/2331-06.
Russell, H. G. (2009) NCHRP Synthesis 393: Adjacent Precast Concrete Box Beam Bridges: Connection Details,
Transportation Research Board of the National Academies, Washington, D.C., 75 pp.
Russell, H. G., Bruce, R. N., and Roller, J. J. (2003) Shear Tests of High Performance Concrete Bulb-Tee Girders.
Proceedings, 3rd International Symposium on High Performance Concrete and PCI National Bridge Conference,
October 19–22, Orlando, FL, Precast/Prestressed Concrete Institute, Chicago, IL, Compact Disc.
Russell, B. W., and Burns, N. H. (1993) Design Guidelines for Transfer, Development and Debonding of Large
Diameter Severn Wire Strands in Pretensioned Concrete Girders. Research Report 1210-5F, Center for Trans-
portation Research, The University of Texas at Austin, 286 pp.
Shahawy, M., Robinson, B., and Batchelor, B.deV. (1993) An Investigation of Shear Strength of Prestressed Concrete
AASHTO Type II Girders. Research Report, Structures Research Center, Florida Department of Transporta-
tion, January 1993.
Shahawy, M. A., and Batchelor, B.deV. (1996) Shear Behavior of Full-Scale Prestressed Concrete Girders: Com-
parison Between AASHTO Specifications and LRFD Code. PCI Journal, 41(3), 48–62.
Tadros, M. K., and Morcous, G. (2011) Impact of 0.7-Inch Diameter Strands on NU I-Girders. Report – Project
No. SPR-1(08) P311, University of Nebraska, pp. 194.
Tadros, M. K., Badie, S. S., and Tuan, C. Y. (2010) NCHRP Report 654: Evaluation and Repair Procedures for
Precast/Prestressed Concrete Girders with Longitudinal Cracking in the Web. Transportation Research Board
of the National Academies, Washington, D.C., 76 pp.
Vecchio, F. J., and Collins, M. P. (1986) The Modified Compression Field Theory for Reinforced Concrete Ele-
ments Subjected to Shear. ACI Journal, 83(2), 219–231.
Wesson, M., (2013) Influence of Strand Debonding on the Shear Strength of Prestressed Concrete. PhD Dissertation,
Purdue University.

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Appendices

The following appendices are not printed herein but are available for download from the
project website (www.trb.org, search for “NCHRP 12-91”):
Appendix A – Survey
Appendix B – Design Case Studies
Appendix C – Finite Element Modeling
Appendix D – Summaries of Individual FEM Simulations
Appendix E – Specimen Details and Fabrication Photographs
Appendix F – Material Properties and Mix Designs
Appendix G – Internal and External Instrumentation
Appendix H – Overview of Design Calculations
Appendix I – AASHTO Bridge Committee Agenda Item
Appendix J – Example of Proposed Changes

103  

Copyright National Academy of Sciences. All rights reserved.


Strand Debonding for Pretensioned Girders

Abbreviations and acronyms used without definitions in TRB publications:


A4A Airlines for America
AAAE American Association of Airport Executives
AASHO American Association of State Highway Officials
AASHTO American Association of State Highway and Transportation Officials
ACI–NA Airports Council International–North America
ACRP Airport Cooperative Research Program
ADA Americans with Disabilities Act
APTA American Public Transportation Association
ASCE American Society of Civil Engineers
ASME American Society of Mechanical Engineers
ASTM American Society for Testing and Materials
ATA American Trucking Associations
CTAA Community Transportation Association of America
CTBSSP Commercial Truck and Bus Safety Synthesis Program
DHS Department of Homeland Security
DOE Department of Energy
EPA Environmental Protection Agency
FAA Federal Aviation Administration
FAST Fixing America’s Surface Transportation Act (2015)
FHWA Federal Highway Administration
FMCSA Federal Motor Carrier Safety Administration
FRA Federal Railroad Administration
FTA Federal Transit Administration
HMCRP Hazardous Materials Cooperative Research Program
IEEE Institute of Electrical and Electronics Engineers
ISTEA Intermodal Surface Transportation Efficiency Act of 1991
ITE Institute of Transportation Engineers
MAP-21 Moving Ahead for Progress in the 21st Century Act (2012)
NASA National Aeronautics and Space Administration
NASAO National Association of State Aviation Officials
NCFRP National Cooperative Freight Research Program
NCHRP National Cooperative Highway Research Program
NHTSA National Highway Traffic Safety Administration
NTSB National Transportation Safety Board
PHMSA Pipeline and Hazardous Materials Safety Administration
RITA Research and Innovative Technology Administration
SAE Society of Automotive Engineers
SAFETEA-LU Safe, Accountable, Flexible, Efficient Transportation Equity Act:
A Legacy for Users (2005)
TCRP Transit Cooperative Research Program
TDC Transit Development Corporation
TEA-21 Transportation Equity Act for the 21st Century (1998)
TRB Transportation Research Board
TSA Transportation Security Administration
U.S.DOT United States Department of Transportation

Copyright National Academy of Sciences. All rights reserved.

You might also like