Sie sind auf Seite 1von 258

VOL. 111, NO.

2
MARCH-APRIL 2014

ACI
STRUCTURAL J O U R N A L

A JOURNAL OF THE AMERICAN CONCRETE INSTITUTE


CONTENTS
Board of Direction ACI Structural Journal
President
Anne M. Ellis March-April 2014, V. 111, No. 2
a journal of the american concrete institute
Vice Presidents an international technical society
William E. Rushing Jr.
Sharon L. Wood
235 
Web Crushing Capacity of High-Strength Concrete Structural
Directors Walls: Experimental Study, by Rigoberto Burgueño, Xuejian Liu, and
Neal S. Anderson Eric M. Hines
Khaled Awad
Roger J. Becker 247 Response of Precast Prestressed Concrete Circular Tanks Retaining
Dean A. Browning Heated Liquids, by Michael J. Minehane and Brian D. O’Rourke
Jeffrey W. Coleman
Robert J. Frosch 257 Bond Strength of Spliced Fiber-Reinforced Polymer Reinforcement,
James R. Harris by Ali Cihan Pay, Erdem Canbay, and Robert J. Frosch
Cecil L. Jones
Cary S. Kopczynski 267 Flexural Behavior and Strength of Reinforced Concrete Beams with
Steven H. Kosmatka Multiple Transverse Openings, by Bengi Aykac, Sabahattin Aykac, Ilker
Kevin A. MacDonald Kalkan, Berk Dundar, and Husnu Can
David M. Suchorski
279 
Experimental Assessment of Inadequately Detailed Reinforced
Past President Board Members Concrete Wall Components, by Adane Gebreyohaness, Charles Clifton,
James K. Wight John Butterworth, and Jason Ingham
Kenneth C. Hover
Florian G. Barth 291 Behavior of Epoxy-Injected Diagonally Cracked Full-Scale Reinforced
Concrete Girders, by Matthew T. Smith, Daniel A. Howell, Mary Ann T.
Executive Vice President Triska, and Christopher Higgins
Ron Burg
303 High-Performance Fiber-Reinforced Concrete Bridge Columns under
Technical Activities Committee Bidirectional Cyclic Loading, by Ady Aviram, Bozidar Stojadinovic, and
Ronald Janowiak, Chair Gustavo J. Parra-Montesinos
Daniel W. Falconer, Staff Liaison 313 Analysis of Early-Age Thermal and Shrinkage Stresses in Reinforced
JoAnn P. Browning
Chiara F. Ferraris
Concrete Walls, by Barbara Klemczak and Agnieszka Knoppik-Wróbel
Catherine E. French 323 Effects of Casting Position and Bar Shape on Bond of Plain Bars, by
Fred R. Goodwin Montserrat Sekulovic MacLean and Lisa R. Feldman
Trey Hamilton
Kevin A. MacDonald 331 Performance of Glass Fiber-Reinforced Polymer-Doweled Jointed
Antonio Nanni Plain Concrete Pavement under Static and Cyclic Loadings, by Brahim
Jan Olek Benmokrane, Ehab A. Ahmed, Mathieu Montaigu, and Denis Thebeau
Michael M. Sprinkel
Pericles C. Stivaros 343 Nonlinear Static Analysis of Flat Slab Floors with Grid Model, by
Andrew W. Taylor Dario Coronelli and Guglielmo Corti
Eldon G. Tipping
353 Effect of Steel Stirrups on Shear Resistance Gain Due to Externally
Staff Bonded Fiber-Reinforced Polymer Strips and Sheets, by Amir Mofidi
Executive Vice President and Omar Chaallal
Ron Burg
363 Punching of Reinforced Concrete Flat Slabs with Double-Headed
Engineering Shear Reinforcement, by Maurício P. Ferreira, Guilherme S. Melo,
Managing Director Paul E. Regan, and Robert L. Vollum
Daniel W. Falconer
375 Behavior of Concentrically Loaded Fiber-Reinforced Polymer Rein-
Managing Editor forced Concrete Columns with Varying Reinforcement Types and
Khaled Nahlawi Ratios, by Hany Tobbi, Ahmed Sabry Farghaly, and Brahim Benmokrane
Staff Engineers
Matthew R. Senecal
Gregory M. Zeisler
Jerzy Z. Zemajtis Contents cont. on next page

Publishing Services
Manager
Discussion is welcomed for all materials published in this issue and will appear ten months from
Barry M. Bergin this journal’s date if the discussion is received within four months of the paper’s print publication.
Discussion of material received after specified dates will be considered individually for publication or
Editors private response. ACI Standards published in ACI Journals for public comment have discussion due
Carl R. Bischof dates printed with the Standard.
Kaitlyn Hinman Annual index published online at http://concrete.org/Publications/ACIStructuralJournal.
Ashley Poirier
ACI Structural Journal
Kelli R. Slayden Copyright © 2014 American Concrete Institute. Printed in the United States of America.
Editorial Assistant The ACI Structural Journal (ISSN 0889-3241) is published bimonthly by the American Concrete Institute. Publica-
tion office: 38800 Country Club Drive, Farmington Hills, MI 48331. Periodicals postage paid at Farmington, MI, and
Tiesha Elam at additional mailing offices. Subscription rates: $161 per year (U.S. and possessions), $170 (elsewhere), payable in
advance. POSTMASTER: Send address changes to: ACI Structural Journal, 38800 Country Club Drive, Farmington
Hills, MI 48331.
Canadian GST: R 1226213149.
Direct correspondence to 38800 Country Club Drive, Farmington Hills, MI 48331. Telephone: 248.848.3700.
Website: http://www.concrete.org.

ACI Structural Journal/March-April 2014 233


387 Repair of Prestressed Concrete Beams with Damaged Steel Tendons Contributions to
Using Post-Tensioned Carbon Fiber-Reinforced Polymer Rods, by ACI Structural Journal
Clayton A. Burningham, Chris P. Pantelides, and Lawrence D. Reaveley
The ACI Structural Journal is an open
397 Study of Composite Behavior of Reinforcement and Concrete in forum on concrete technology and papers
Tension, by John P. Forth and Andrew W. Beeby related to this field are always welcome.
407 Flexural Capacity of Fiber-Reinforced Polymer Strengthened Unbonded All material submitted for possible publi-
cation must meet the requirements of
Post-Tensioned Members, by Fatima El Meski and Mohamed Harajli
the “American Concrete Institute Publi-
419 Size Effect on Strand Bond and Concrete Strains at Prestress Transfer, cation Policy” and “Author Guidelines
by José R. Martí-Vargas, Libardo A. Caro, and Pedro Serna-Ros and Submission Procedures.” Prospective
authors should request a copy of the Policy
431 Proposed Minimum Steel Provisions for Prestressed and Nonpre- and Guidelines from ACI or visit ACI’s
stressed Reinforced Sections, by Natassia R. Brenkus and H. R. Hamilton website at www.concrete.org prior to
441 Lateral Strain Model for Concrete under Compression, by Ali Khajeh submitting contributions.
Samani and Mario M. Attard Papers reporting research must include
a statement indicating the significance of
453 Discussion the research.

Cyclic Loading Test for Beam-Column Connection with Prefabricated Reinforcing The Institute reserves the right to return,
Bar Details. Paper by Tae-Sung Eom, Jin-Aha Song, Hong-Gun Park, Hyoung-Seop without review, contributions not meeting
Kim, and Chang-Nam Lee the requirements of the Publication Policy.
All materials conforming to the Policy
Shear Strength of Reinforced Concrete Walls for Seismic Design of Low-Rise requirements will be reviewed for editorial
Housing. Paper by Julian Carrillo and Sergio M. Alcocer quality and technical content, and every
effort will be made to put all acceptable
Performance of AASHTO-Type Bridge Model Prestressed with Carbon Fiber-
papers into the information channel.
Reinforced Polymer Reinforcement. Paper by Nabil Grace, Kenichi Ushijima, However, potentially good papers may be
Vasant Matsagar, and Chenglin Wu returned to authors when it is not possible
460 In ACI Materials Journal to publish them in a reasonable time.

462 Reviewers in 2013 Discussion


All technical material appearing in the
MEETINGS ACI Structural Journal may be discussed.
If the deadline indicated on the contents
page is observed, discussion can appear
2014 27—The Changing Future of Cement in the designated issue. Discussion should
& Concrete: Threat or Opportunity, be complete and ready for publication,
MARCH Leicestershire, United Kingdom, http://ict. including finished, reproducible illustra-
concrete.org.uk tions. Discussion must be confined to the
19-21—ICRI 2014 Spring Convention, scope of the paper and meet the ACI Publi-
Reno, NV, www.icri.org/Events/2014_ MARCH/APRIL cation Policy.
Spring/conv_home.asp Follow the style of the current issue.
30-2—ACPA 2014 Convention, Be brief—1800 words of double spaced,
22—ASA Spring 2014 Committee Indianapolis, IN, http://convention.myacpa. typewritten copy, including illustrations
Meetings, Reno, NV, www.shotcrete.org org/indy2014/ and tables, is maximum. Count illustrations
and tables as 300 words each and submit
24-29—ICPI Annual Meeting, New them on individual sheets. As an approxi-
Orleans, LA, www.icpi.org/node/3996 mation, 1 page of text is about 300 words.
Submit one original typescript on 8-1/2 x
11 plain white paper, use 1 in. margins,
UPCOMING ACI CONVENTIONS and include two good quality copies of the
The following is a list of scheduled ACI conventions: entire discussion. References should be
complete. Do not repeat references cited
2014—March 23-27, Grand Sierra Resort, Reno, NV in original paper; cite them by original
2014—October 26-30, Hilton Washington, Washington, DC number. Closures responding to a single
2015—April 12-15, Marriott & Kansas City Convention Center, Kansas City, MO discussion should not exceed 1800-word
equivalents in length, and to multiple
For additional information, contact: discussions, approximately one half of
Event Services, ACI the combined lengths of all discussions.
38800 Country Club Drive, Farmington Hills, MI 48331 Closures are published together with
Telephone: 248.848.3795 the discussions.
e-mail: conventions@concrete.org Discuss the paper, not some new or
outside work on the same subject. Use
references wherever possible instead of
ON COVER: 111-S23, p. 269, Fig. 3—Diagonal reinforcement spiraling around circular openings. repeating available information.
Discussion offered for publication should
offer some benefit to the general reader.
Discussion which does not meet this
Permission is granted by the American Concrete Institute for libraries and other users registered with the Copyright requirement will be returned or referred to
Clearance Center (CCC) to photocopy any article contained herein for a fee of $3.00 per copy of the article. Payments
should be sent directly to the Copyright Clearance Center, 21 Congress Street, Salem, MA 01970. ISSN 0889-3241/98 the author for private reply.
$3.00. Copying done for other than personal or internal reference use without the express written permission of the
American Concrete Institute is prohibited. Requests for special permission or bulk copying should be addressed to the
Managing Editor, ACI Structural Journal, American Concrete Institute. Send manuscripts to:
The Institute is not responsible for statements or opinions expressed in its publications. Institute publications are not able http://mc.manuscriptcentral.com/aci
to, nor intend to, supplant individual training, responsibility, or judgment of the user, or the supplier, of the information
presented.
Send discussions to:
Papers appearing in the ACI Structural Journal are reviewed according to the Institute’s Publication Policy by individuals
expert in the subject area of the papers. Journals.Manuscripts@concrete.org

234 ACI Structural Journal/March-April 2014


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 111-S04
Note: Paper 111-S04 of the January-February 2014 ACI Structural Journal has been reprinted herein with corrected figure art.
Please reference the March-April 2014 version of this paper only.

Web Crushing Capacity of High-Strength Concrete


Structural Walls: Experimental Study
by Rigoberto Burgueño, Xuejian Liu, and Eric M. Hines
This paper discusses the relationship between concrete strength possibility because they have observed deformed configura-
and web crushing capacity based on results from large-scale tests tions in the inelastic range that could undermine the benefits
of thin-webbed structural walls with confined boundary elements. of increased concrete strength. The work described herein
Eight walls with concrete strengths ranging from 39 to 138 MPa represents an attempt to answer this question experimentally
(5.6 to 20 ksi) were tested to web crushing failure under cyclic
and thereby establish limits for the future analytical prediction
and monotonic loading. These tests clearly demonstrated differ-
of web crushing failures. This experimental program proceeded
ences between elastic and inelastic web crushing behavior and
their dependence on concrete strength. Walls with higher concrete with the intention of testing the following two hypotheses:
strengths reached higher levels of displacement ductility due to an 1. Web crushing strength increases in proportion to fc′ as long
increase in web crushing capacity. Evidence with respect to mono- as the struts are not damaged. Hence, transformation from an
tonic tests showed that degradation of the diagonal compression elastic web crushing failure to an inelastic web crushing failure
struts from cyclic loading increases with concrete strength, thus can be achieved simply by increasing the concrete strength; and
limiting the inelastic deformation capacity gains. Thus, concrete 2. Damage to struts caused by cyclic loading and inelastic
compressive strength does not linearly increase web crushing deformations can limit web crushing strength independently
strength as implied by rational web crushing models; rather, the of fc′. Hence, increases in fc′ may not lead to proportional
relationship is nonlinear, with a decreasing limit as concrete increases in ductility capacity.
strength increases. The ACI shear stress limit considerably under-
The experimental program designed to test these two
estimated the web crushing capacity of the walls. Test results and
hypotheses consisted of two sets of four structural walls with
observations are reported with the intent of providing physical
insight into the web crushing failure mechanism and the inherent a range of concrete strengths. One set of walls was tested
limits of thin-webbed concrete members in shear. cyclically, and the other set was tested monotonically.

Keywords: ductility; high strength; shear walls; web crushing. RESEARCH SIGNIFICANCE
Elastic and inelastic web crushing failures were consis-
INTRODUCTION tently achieved in a series of large-scale structural wall
Over the past 50 years, web crushing capacity of reinforced tests designed to study the relationship between concrete
concrete members has emerged as a primary design concern in compressive strength and web crushing strength of thin-
three distinct contexts: gravity loading of thin-webbed beams in webbed members. The results of these tests validate both the
the 1960s,1-4 seismic loading of structural walls with confined dependence of web crushing capacity on fc′ and the signif-
boundary elements in the 1970s,5-9 and seismic loading of icant degradation of web crushing capacity experienced for
hollow bridge piers with confined corner elements in the a range of concrete strengths under cyclic loading in the
1990s.10-13 In each context, motivation to design lightweight inelastic range. Physical insight developed from observa-
members based on physical insight led to large-scale struc- tions and measurements of these tests provides a firm foun-
tural testing programs that discovered web crushing capaci- dation for establishing the limits of thin-webbed reinforced
ties significantly in excess of the average shear stress limits concrete member design. Consistency of the test results indi-
recommended by ACI-318.14 Researchers in charge of these cates that it may be possible to design thin-webbed elements
testing programs have repeatedly emphasized the importance to experience significant inelastic deformations before
of understanding shear behavior in terms of diagonal tension failing in shear, opening up new possibilities for acceptable
and diagonal compression instead of average shear stresses. ductile failure modes of reinforced concrete members.
While diagonal compression demands depend on member
geometry and reinforcement, diagonal compression capacities WEB CRUSHING SHEAR CAPACITY
depend on the size and strength of the most heavily loaded Work in the 1960s on thin-webbed girders established
struts. Previous research programs established consensus that properly reinforced concrete webs could resist diag-
regarding the linear dependence of web crushing capacity on onal compression loads in the elastic range almost equal
concrete compressive strength fc′ and web thickness. Limits
ACI Structural Journal, V. 111, No. 2, March-April 2014.
on the value of fc′ itself, however, were not evaluated. Could MS No. 2011-322.R2, doi:10.14359.51686515, was received May 23, 2013, and
the web crushing capacity of a 30 MPa wall be increased by reviewed under Institute publication policies. Copyright © 2014, American Concrete
Institute. All rights reserved, including the making of copies unless permission is
a factor of four simply by increasing the concrete strength to obtained from the copyright proprietors. Pertinent discussion including author’s
closure, if any, will be published ten months from this journal’s date if the discussion
120 MPa? Seismic researchers have hesitated to endorse this is received within four months of the paper’s print publication.

ACI Structural Journal/March-April 2014 235


Fig. 1—Elastic and inelastic shear web crushing: (a) definition of elastic and inelastic web crushing failure modes and critical
regions in wall; (b) free body diagrams used for assessing flexure-shear web crushing capacity of structural wall with confined
boundary elements13; and (c) schematic of strut degradation due to cyclic loading.
to the compressive strength of concrete itself, resulting in on individual struts inside the plastic hinge region as it
average shear stresses an order of magnitude higher than spreads up the height of a wall. As shown in Fig. 1(a), two
the prevailing provisions.15 It was during this work that the distinct shear transfer mechanisms were identified for the
term “web crushing” was conceived, and a case was made plastic hinge region and the regions elsewhere. Elastic, or
to consider diagonal compression capacity independently of standard shear, struts are formed in the wall web in regions
average shear stresses. Work in the 1970s, 1980s, and early that have not experienced significant tensile strains along
1990s on structural walls with confined boundary elements both the longitudinal (vertical) and transverse directions,
under seismic loads emphasized the dependence of web leading to a parallel shear cracking pattern (at an angle θs).
crushing capacity on inelastic deformation and the realign- In other words, this region is stressed mainly under in-plane
ment of cracks in the web crushing region, but recommended shear stress while the effect of elastic flexural strain is not
assessment of web crushing capacity in terms of average significant. By comparison, plastic flexural strains force the
shear stresses.9,16 Among these works, the effect of concrete struts inside the plastic hinge region to realign so that they
strength on web crushing capacity was witnessed through all converge in the flexural compression toe. This can be
the tests by the Portland Cement Association (PCA)7,9 in the understood by considering that the flexural crack at the base
mid-1970s on walls with boundary elements. Wall B6 with of the wall prohibits shear force transfer into the footing at
a concrete compressive strength of 22 MPa (3165 psi) failed any location except for the flexural compression toe, and that
in web crushing at significantly lower deformation capacity the struts should fan upward until they are able to carry the
than Wall B7 with a concrete compressive strength of full elastic shear force. These fanning struts are considered
49 MPa (7155 psi). No further high-strength concrete (HSC) as inelastic or flexure-shear struts.
structural walls were tested, however. Work in the late 1990s Based on the noted force transfer mechanisms, Hines and
and early 2000s on hollow bridge piers with confined corner Seible13 proposed a web crushing capacity model through
elements clearly distinguished between elastic web crushing equilibrium analysis of the free body diagrams of isolated
and inelastic web crushing, and tied the assessment of elastic (or standard) and inelastic (or flexure-shear) diagonal
inelastic web crushing to the development of the plastic hinge struts (Fig. 1(b)). Calculating the demand of forces on the
region,12 emphasizing the interaction of inelastic flexure and elastic struts NDs and the compressive capacity NCs of these
shear behavior in this zone.13 The physically based method struts leads to the standard shear web crushing equation
of assessment by Hines and Seible13 recognized the focus proposed by Oesterle et al.7 and Paulay and Priestley16
on principal stresses from the 1960s research, accounted for
inelastic deformations, and allowed accurate assessment of NCs
≤ kfc′sin θ s cos θ s (1)
cross sections with various relations between depth of web N Ds
and depth of boundary elements. A review of the model is
provided as follows; however, for a detailed description of
the model, the reader should refer to Reference 13. The inelastic struts in the plastic hinge region differ from
The approach to web crushing capacity by Hines and their elastic counterparts both in demand and capacity. On
Seible13 is based on the assessment of capacity and demand the demand side, the inelastic struts are required to transfer
inelastic shears at angles that are consistent with the spread

236 ACI Structural Journal/March-April 2014


of plasticity Lpr. On the capacity side, these struts become
narrower toward the compression toe from which they fan.
Following Hines and Seible,13 the critical flexure shear
crack angle θfs is calculated as

 A f  
 jd  v yv + f1t w  
  s    jd 
θ fs = tan −1  −1
 = tan  L  (2)
 (
2 T − Tyav )   pr 
 
 

Web crushing is then estimated to occur when the capacity


NCfs of the critical strut (Eq. (3)) is equal to the demand NDfs
on this strut (Eq. (4))
Fig. 2—Analytical force-displacement response with web
NCfs = kfc′t w Rdθ (3) crushing capacity predictions.

of yielding, while a 138 MPa (20 ksi) wall would fail in
flexure. The model, however, assumes integrity of the struts,
∆T which are known to degrade with increase ductility demand
N Dfs = − f1 st w sin θ fs (4) and cyclic loading as schematically shown by the cracking
cos θ fs
pattern in Fig. 1(c). Nonetheless, the suggestion that web
crushing strength can be directly and consistently related to
The related variables are determined from a moment-cur- concrete compressive strength is indicative of new possibili-
vature analysis of the cross section and the strut geometry, ties for increasing shear capacities of thinned-webbed struc-
as shown in Fig. 1(b). The concrete compressive strength tural members with increased concrete strength.
softening factor k is calculated according to the modified
compression field theory17 with a simplified approach for EXPERIMENTAL PROGRAM
determining the principal tensile strain ε1. It should be noted Test unit identification, geometry, and
that while the model has no explicit limit on fc′, its predic- reinforcement details
tion quality for HSC is uncertain because the available test To verify the aforementioned hypotheses and establish
data for calibration of the model was from tests of normal- rational performance levels on the inelastic web crushing
strength concrete (NSC) walls. limits for HSC structural walls, eight 1/5-scale cantilever
ACI 318-1114 does not reflect the direct dependence of web structural walls with design concrete compressive strengths
crushing capacity on fc′, but rather defines the expression in of 34, 69, 103, and 138 MPa (5, 10, 15, and 20 ksi) were
Eq. (1) as 0.83√fc′ (MPa) (10√fc′ [psi]), although √fc′ is a tested under cyclic and monotonic loading.18 The walls
quantity commonly related to concrete tensile strength and had a barbell-type cross section with thin webs and heavily
diagonal tension failures. A maximum value of 0.69 MPa confined boundary elements, and were designed to induce a
(100 psi) is adopted by the code because of the lack of test web crushing failure and not to represent a component from
data and practical experience with concrete strengths more a prototype structure. The test unit cross sections with rein-
than 69 MPa (10,000 psi). forcement details are shown in Fig. 3. The identification
The proportional increase of web crushing capacity with name for the walls starts with M, followed by two digits
fc′ implied by Eq. (3) indicates the potential of achieving denoting the design concrete compressive strength in kip/in.2
a ductile force-displacement response in structural walls (1 kip/in.2 [ksi] = 6.895 MPa) and then by a letter describing
even if ultimately limited by web crushing. Figure 2 shows the loading protocol: C for cyclic and M for monotonic
simulated force-displacement responses for structural walls loading. For example, test unit M10C refers to the wall with a
with heavily reinforced boundary elements12 along with design concrete strength of 69 MPa (10 ksi) and subjected to
web crushing predictions. Curves for walls with concrete cyclic loading. All walls had an effective length of 2540 mm
compressive strengths varying from 34 to 138 MPa (5 to (100 in.) for an aspect ratio (M/V) of 2.5. In all cases, the
20 ksi) are shown. The web-crushing capacity envelopes wall web was 508 mm (20 in.) deep and 76 mm (3 in.) thick.
after Hines and Seible13 and the ACI Code14 limits are also The boundary elements had a depth of 254 mm (10 in.). As
plotted. It can be seen that force-deformation response of shown in Fig. 3 and Table 1, the steel reinforcement was
the walls is only slightly affected by the increased concrete essentially the same for all walls, with a small variation in
compressive strength. The Hines and Seible model, however, the longitudinal reinforcement of the boundary elements in
predict the web crushing capacity to increase dramatically Wall M15C and in the transverse reinforcement spacing for
with an increase of concrete strength. The ACI limit is Walls M20C and M20M. The reinforcement ratios for the
clearly conservative and independent of the inelastic defor- web and boundary elements of the test units are given in
mations in the wall. The Hines and Seible model implies that Table 2.
a 34 MPa (5 ksi) wall would fail by web crushing at the onset

ACI Structural Journal/March-April 2014 237


Table 1—Test unit steel reinforcement material
properties and layout
Test Spacing, fy, fu,
unit Bar Size mm MPa MPa εsh Esh, MPa
M1 M25 4 bars 524 —* — 11,034
M2 M22 4 bars 448 672 0.0026 11,262
M05C
m M10 127 445 692 0.0083 7759
M05M
T M10 102 445 692 0.0083 7759
S M10 76 459 703 0.0075 7759
M1 M25 4 bars 464 697 0.0096 8966
M2 M22 4 bars 448 672 0.0094 8966
M10C
m M10 127 476 746 0.0060 8966
M10M
T M10 102 476 746 0.0060 8966
S M10 76 545 730 0.0050 7586
M M19 12 bars 439 705 0.0078 8966
m M10 127 481 759 0.0054 9655
M15C
T M10 102 481 759 0.0054 9655
S M10 102 503 756 0.0030 7931
Fig. 3—Test unit cross sections with reinforcement details
(refer also to Table 3).
M1 M25 4 bars 586 — — 2759
M2 M22 4 bars 421 630 0.0093 8621 Loading protocol
The test setup (Fig. 4) was designed to load the walls
M15M m M10 127 478 748 0.0060 10,690
in-plane as cantilevers. The walls were loaded monotoni-
T M10 102 478 748 0.0060 10,690 cally and cyclically with constant axial load. The axial
S M10 76 510 656 0.0072 5517 load for all test units was 579 kN (130 kip), corresponding
M1 M25 4 bars 451 703 0.0054 9310 to 0.10fc′Ag for a reference concrete strength of 34 MPa (5
ksi). Axial load was applied using hydraulic jacks and high-
M2 M22 4 bars 446 699 0.0060 8966
M20C* strength rods reacting against the top load stub through a
m M10 127 438 703 0.0043 8793 spandrel beam. The horizontal load was applied with a
M20M
T M10 *
76 , 102 438 703 0.0043 8793 servo-controlled actuator connected to a load stub at the
S M10 76 443 717 0.0037 8621 top of the wall. Lateral stability was provided by a pair of
*
parallel inclined tensioned chains on both sides of the wall.
— is not displayed in response.
Cyclic and monotonic loading, respectively, were applied
Note: 1 MPa = 0.145 ksi.
on two test units with the same design concrete strength
Table 2—Test unit steel reinforcement ratios to assess shear strength/stiffness degradation and inelastic
web crushing limits under different loading histories. Cyclic
Test unit ρl ρn ρs ρh tests were done according to an incrementally increasing
M15C 0.0528 0.0147 0.0357 0.0183 fully reversed cyclic pattern. Four cycles (in quarter incre-
M20C and M20M 0.0556 0.0147 0.0237 0.0244 ments) were first applied in force control until the theoret-
ical first yield force, Fy′, defined as the force at the onset of
All others 0.0556 0.0147 0.0237 0.0244
yield of the extreme longitudinal reinforcing bar in tension
as obtained from a moment-curvature analysis. The top
Material properties
displacement at the theoretical first yield force, Δy′, deter-
The HSC for this research was attained with traditional
mined by the average of the measured values from the posi-
mixture constituents following examples from commer-
tive and negative loading excursions, was used to define
cially available mixtures.18 Compressive, tensile, and flex-
the experimental elastic bending stiffness, KE = Fy′/Δy′. The
ural strengths were assessed through standard testing, and
ideal yield displacement,19 Δy, corresponding to displace-
the values at the day of testing for all walls are provided in
ment ductility one (μΔ = 1), was determined using the exper-
Table 3. Table 1 lists the properties for the steel reinforce-
imental stiffness at first yield and the ideal yield force, Fy,
ment with reference to the nomenclature noted in Fig. 3. The
by Δy = Fy/KE. The ideal yield force Fy was computed by
properties given in Table 1 are average values from three
means of a moment-curvature analysis and corresponded to
tensile tests on 457 mm (18 in.) long segments for each of
the moment at the critical section at which either the extreme
the reinforcement bars.
confined concrete fibers reached εc = 0.004 or the extreme
steel fiber in tension reached εs = 0.015, whichever occurred
first.19 The remainder of the test was conducted in displace-
ment control with two cycles each at the system displace-

238 ACI Structural Journal/March-April 2014


Table 3—Test unit concrete material properties

Test Design fc′, MPa ft′, MPa fr′, MPa


unit fc′, MPa x σ x σ x σ
M05C 34 46.0 1.37 3.25 0.207 5.49 0.0897
M05M 34 38.9 1.23 3.55 0.0966 5.37 0.0551
M10C 69 56.4 1.86 4.50 0.331 6.57 0.221
M10M 69 84.0 1.37 5.54 0.490 7.33 0.669
M15C 103 102 1.01 5.70 0.441 9.01 0.559
M15M 103 111 4.99 6.17 0.910 0.935 1.08
M20C 138 131 3.01 6.19 0.400 11.5 0.359
M20M 138 115 2.55 5.96 0.172 10.2 0.593

Note: 1 MPa = 0.145 ksi.

Table 4—Force and displacement values at


theoretical and ideal yield
Test M05C M05M M10C M10M M15C M15M M20C M20M
Fy′,
578 576 583 618 586 644 576 572
kN
Δy′,
17.6 18.5 18.4 15.8 15.7 16.2 13.5 15.2
mm
Fy,
842 836 723 745 731 816 809 803
kN
Δy,
25.7 26.9 22.9 19.1 19.6 20.6 19.1 21.3
mm
Note: 1 kN = 0.225 kip; 1 mm = 0.0394 in.

shear deformations were measured on two wall segments


(one from the wall base to 1016 mm [40 in.] high, and
the second one from the noted level to the wall top) with
a pair of displacement transducers arranged in opposing
diagonal directions on one wall face. Global displacement
at the effective height of the wall (2540 mm [100 in.]) was
measured with a string potentiometer. Further details of the
instrumentation are reported elsewhere.18

OBSERVATIONS
Common behavior
Common behavior observed on all test units is described
herein. No cracking was observed up to 0.25Fy′. At 0.5Fy′,
flexural cracking in the boundary elements and diagonal
Fig. 4—Test setup overview. shear cracking in the webs appeared. At 0.75Fy′, elastic shear
cracking developed throughout the entire web. The HSC
ment ductility levels μΔ = 1, 1.5, 2, 3, 4, and 6, or until walls developed a denser pattern of flexural and shear cracks
failure of the test unit. The monotonic tests were conducted than the NSC walls. With increasing displacement ductility
by applying the lateral load in force control until Fy′, and crack density increased and the active cracks became wider.
then in displacement control until failure. The values of Fy′ Overall, shear crack spacing in the HSC walls was much
and Δy that defined the loading protocol are listed in Table 4. smaller than for the NSC walls. Even though diagonal shear
cracking under tension does not control the capacity of the
Instrumentation walls, it defines the height and width of the struts, and thus
The walls were instrumented to measure segmental flex- affects the strut capacities. Strut capacity is also affected by
ural curvatures, shear deformations, and steel reinforcement crack width, which dictates shear slip behavior at the crack
strains. The layout of the external instrumentation is shown interface. Finally, relatively larger cover concrete spalling
in Fig. 5. Flexural section curvatures were calculated using on the compression boundary element was observed on the
displacement transducers placed along the height of the HSC walls. Nonetheless, the heavily confined boundary
column on both sides of the boundary elements. Average elements had no problem resisting the compression force.

ACI Structural Journal/March-April 2014 239


crack spacing was much smaller. The fanning flexure-shear
cracking pattern was formed within the plastic hinge region,
with fairly flat cracks close to the bottom and much steeper
cracks at the top. An example of this inelastic flexure-shear
failure mode is shown in Fig. 6(d) for Wall M10M.
Loading protocol had a large influence on damage and
failure patterns. This can be seen by comparing the failure
modes between the cyclic and monotonic tests for Walls M15
and M20 (Fig. 6(e) to (h)). The monotonically loaded units
developed a denser crack pattern, with multiple branching
and wedge-shaped struts. Conversely, the cyclic tested walls
had wider spaced cracks and struts with a more uniform
width. The uniform damage pattern from denser cracking
in the monotonic tests allowed for more diagonal compres-
sion load paths in the wall web. The consistent loading also
permitted the struts to remain integral, thus allowing these
walls to sustain larger inelastic deformations. In constrast,
the wider cracking pattern from cyclic loading led to larger
crack misalignment and damage to the diagonal struts, which
reduced deformation capacity.
The cracking pattern due to cyclic loading is illustrated in
Fig. 1(c) and shown in Fig. 7, which illustrates the degra-
Fig. 5—External instrumentation layout (units in mm). dation of inelastic struts for wall M20C upon reaching the
(Note: 1 mm = 0.0394 in.) second negative displacement target (second cycle) for
μΔ = 4. Figure 7(a) is a view of the bottom region of the
Damage patterns and failure mechanisms
wall from which the fanning crack pattern inside the plastic
Peculiarities on damage pattern, web crushing failure,
zone can be discerned. The close-up view of the wall web
and ductility capacity are described herein individually. All
in Fig. 7(b) shows the crisscross cracking pattern from the
walls failed in web crushing according to the experimental
reversed loading cycles. Upon reloading, crack misalignment
aim. Walls M05C and M10C failed on the first excursion to
results from shear deformations due to yielding and bond-slip
μΔ = 2, and web crushing occurred at μΔ = 1.8. In both walls,
effects in the longitudinal and transverse web reinforcement.
there was a sudden loss of strength upon failure of approxi-
As cracks close for the compression load path to reestablish in
mately 40%. Wall M15C performed in a ductile manner up
the previously formed struts, the crack misalignment induces
to μΔ = 4, and failed in web crushing on the second excursion
large local stresses due to shear friction and distortion of the
to ductility 4. This HSC wall had a gradual loss of strength
struts. This causes the web cover concrete to lose its bond
upon failure, losing only approximately 8% of its strength
to the reinforcement and spall off, as seen in the lower left
compared with the load reached during the first cycle at
region of the wall web (Fig. 7(b)). The test units gradually
ductility 4. Wall M20C performed in a ductile manner up
lost their load-carrying capacity as a result of the diminished
to its failure by web crushing on the first excursion to μΔ =
load transfer efficiency of the concrete struts. Web crushing
6. Web crushing started to develop at a top displacement of
in the cyclically loaded walls was observed to expand over a
66 mm (2.6 in.), and the wall strength degraded by roughly
large area within the plastic hinge region, and crushing of the
40% when the displacement reached the target displacement
flexure-shear struts initiated in the center of the web and then
for ductility six. Wall M05M failed by elastic web crushing
extended to the edge of the compression boundary element
at μΔ = 2.3. Walls M10M, M15M, and M20M performed in
(Fig. 6(e) and (g)). The noted differences between the cyclic
a ductile manner up to web crushing failure at μΔ = 7, 6.5,
and monotonic responses became more significant for higher
and 9.2, respectively.
values of concrete compressive strength.
The damage and failure patterns in the wall bottom third
region (850 mm [33 in.]) for all test units are shown in Fig. 6.
RESULTS
Test units with lower concrete compressive strength (M05C,
Force-deformation response
M05M, and M10C) failed after only minor levels of inelastic
The hysteretic force-displacement response of the four
response (Fig. 6(a) to (c)). Tensile cracking was minimal, and
walls under cyclic loading is shown in Fig. 8. Again, failure
cracks fully closed upon load reversal. No crack realignment
in all cases was due to web crushing. Recalling that the rein-
was observed and thus these walls were limited by standard,
forcement details were essentially the same for all walls, the
or elastic, web crushing. The failures were sudden and the
test results demonstrate that increased concrete compressive
crushing of the concrete struts occurred along the interface
strength allowed the walls to considerably increase their
of the wall web and the compression boundary element, and
inelastic deformation capacity by delaying shear failure. Web
instantly propagated along the wall height. The rest of the
crushing was thus shifted from an essentially elastic level for
test units exhibited moderate to high ductile behavior before
Wall M05C, failing after completion of two cycles at μΔ = 1.5,
web crushing failure. Cracking was more extensive, and
to a highly stable ductile response for Wall M20C, failing after

240 ACI Structural Journal/March-April 2014


Fig. 6—Damage and failure patterns in wall bottom third region (850 mm [33.5 in.]).
sustaining two full cycles at μΔ = 4. Comparison of inelastic idealized bilinear moment-curvature response.19 The curva-
deformation capacity in terms of displacement ductility is ture profiles of the M10 and M20 walls are shown in Fig. 9.
adequate because the ideal yield displacement for all walls For cyclic loaded test units that failed at low to medium
did not vary greatly, as shown in Table 4. The ductile behavior ductility levels, the curvature distribution along the height
displayed by Walls M15C and M20C shows that these units was almost linear until failure. For monotonic loaded test
preserved the high stiffness and lateral load-carrying capacity units that failed at a high ductility level, the plastic rotation
characteristic of structural walls, mostly provided by the web, was mainly concentrated within 300 mm (12 in.) from the
while benefiting from the inelastic deformation capacity of bottom of the wall. It can be seen, however, that the spread
column flexural hinges at the boundary elements. Finally, of plasticity is very similar at equal displacement ductility
the energy dissipation capacity, as judged by the area of the levels for both monotonic and cyclic loaded walls. The curva-
hysteresis loops, is also notably increased solely due to the ture distributions provide local-level evidence of the signif-
increase of the concrete compressive strength. icant inelastic flexural deformations sustained by the walls.
Furthermore, the observed fanning crack pattern (Fig. 6) and
Inelastic behavior characteristics the essentially linear distribution of plastic curvatures along
According to Eq. (2) the length of the plastic hinge region the plastic hinge region (with the linearity only disturbed by
Lpr is directly proportional to the angle of the flexural-shear boundary effects at the footing) suggests that Eq. (4) can be
cracking, and hence the force demand on the critical inelastic used to assess the demand on the critical inelastic strut.
strut. The spread of plasticity can be taken as the length over Table 5 shows the separate contribution of flexure and
which plastic curvatures exceed the yield curvature from an shear effects to the displacement at the wall top for the cycli-

ACI Structural Journal/March-April 2014 241


Fig. 7—Degradation of inelastic struts for M20C wall during ductility 4 cycling: (a) view of wall bottom third region (850 mm
[33.5 in.]); and (b) close-up view of wall web.

Fig. 8—Hysteretic force-displacement response of cyclic tests. (Note: 1 MPa = 0.145 ksi; 1 mm = 0.0394 in.)
cally tested units at the first positive peak of each ductility web. Shown in Fig. 11(c) are longitudinal strain profiles for the
level. The flexure-induced displacement Δf at the top of M20 walls, which have a linear variation along the height (with
the wall was calculated by adding the contribution of indi- disturbance near the footing). The strains were calculated using
vidual sectional rotations. The top wall displacement due to the displacement transducers along both sides of the boundary
shear Δs was calculated by summing the shear deformations elements (Fig. 5). The profiles are consistent with the moment
measured on the two wall segments (Fig. 5). From the data gradient on the wall. It can be further observed that the strain
in Table 5 it can be confirmed that the shear displacements profiles for both M20 walls are essentially the same at equal
are linearly related to the flexural displacements.12 For the displacement ductility demands. This was consistent for the
eight tested walls,18 the average ratio of shear to flexural other three wall sets, supporting the evidence of equal flexural
displacements was 0.23, with a standard deviation of 0.04. demands on monotonic and cyclic tests.
Figure 10 shows the average shear stress versus shear strain Principal strains in the wall web were estimated from consid-
hysteretic response of Wall M20C in the bottom and top wall eration of a whole wall segment (web and boundary elements)
segments. The shear deformations were mainly concentrated through Mohr’s circle with the measured longitudinal strains,
in the bottom wall segment (1016 mm [40 in.] from the base), the measured average shear strains (Fig. 5), and neglecting the
which is where the plastic hinge region develops. It can be transverse strains (due to the presence of the heavily reinforced
seen that the average shear stresses considerably exceeded boundary elements). Principal strain values calculated this way
the ACI limits. This observation applies to all walls because at the web mid-depth are shown in Fig. 11(b) against displace-
they had similar levels of lateral load resistance. ment ductility for all walls. Except for some deviations for
To provide further insight and quantitative information on the M05 walls, the average principal strains were equal for all
the performance of the tested walls, Fig. 11 provides a brief walls at the same displacement ductility level.
overview of average strain demands at mid-depth on the wall

242 ACI Structural Journal/March-April 2014


Fig. 9—Average curvature strain profiles for: (a) M10 walls; and (b) M20 walls.

Fig. 10—Average shear stress-strain response of: (a) bottom; and (b) top segments of Wall M20C.
Monotonic versus cyclic loading Table 5—Flexure and shear components of wall
A comparison of the force-displacement envelopes of top displacement at first positive peak of each
the cyclic and monotonic tests is shown in Fig. 12. It is ductility level
clear that higher fc′ resulted in higher inelastic deformation Ductility Displacement M05C M10C M15C M20C
capacity. The force-deformation response of the walls, up
Δf, mm 18.5 18.0 14.5 14.5
to their respective deformation limit, is considered to have μΔ = 1.0
been essentially the same, with minor differences due to: a) Δs, mm 5.08 6.10 3.53 3.30
variations in fc′; b) longitudinal reinforcement differences Δf, mm 28.4 27.7 22.6 23.1
for Wall M15C (Fig. 3); c) earlier spalling in the compres- μΔ = 1.5
Δs, mm 7.62 7.87 4.83 4.57
sion toe for Wall M20C; and d) reduction of the effective
Δf, mm — — 30.5 30.7
concrete compressive strength due to more severe cracking μΔ = 2.0
for the cyclically loaded walls. Δs, mm — — 6.60 6.60
The increase in deformation capacity, however, was not Δf, mm — — 47.0 47.0
directly proportional to fc′, particularly for the cyclic loaded μΔ = 3.0
Δs, mm — — 9.40 9.91
walls. The monotonic and cyclic deformation capacities of
Δf, mm — — 62.5 62.0
the M05 walls were essentially the same (Fig. 12(a)). The μΔ = 4
response was similar because both walls failed close to Δs, mm — — 13.2 15.5
the elastic range, and only minimal cycling was done on Note: 1 mm = 0.0394 in.
Wall M05C. The responses of the other walls show that
while increased concrete strength leads to larger deforma- significant effect of cyclic loading on the deformation limit
tion capacity, cyclic loading curtails this improvement. of Wall M20C despite the larger concrete strength in M20C
The deformation capacity reduction of the cyclically tested compared with that in M20M. It is thus hypothesized that the
walls is attributed to the damage of the flexure-shear struts reduced deformation capacity in the cyclically tested HSC
from cyclic loading, which reduces their capacity to transfer walls is due to the negative convergence of an increased
load from the tension to the compression boundary element. stress intensity field at crack misalignment and a reduced
This effect is best seen by observing the responses for the M15 cracking bridge zone from the higher strength concrete. This
walls in Fig. 12(c). The effect was not as clearly captured for can be understood upon considering that HSC experiences
the M10 walls (Fig. 12(b)) because fc′ for M10C was lower dramatic strength degradation in the postpeak response, and
than that for M10M. Nonetheless, given the response of Wall the effect of further strength increase is not appreciable. For
M15C, it can be expected that if fc′ for Wall M10C had been structural walls, the later effect would indicate a curtailing
closer to the design target, its deformation capacity would effect on the increased inelastic deformation gains on web
have been increased, and the cyclic and monotonic envelopes crushing capacity for increasing values of fc′.
would have been similar to those obtained for the M15 walls. The reduced gain in deformation capacity of Wall M20C
Comparison of the M20 wall response envelopes indicates a compared with that of Wall M15C would seem to indicate

ACI Structural Journal/March-April 2014 243


crushing, while the rest of the walls, which failed by inelastic
web crushing, followed Eq. (3). Figure 13(a) plots the calculated
values of k versus normalized shear distortions in the plastic
hinge region. The maximum (condition before web crushing)
average shear distortion within the plastic hinge region, γm, was
normalized by the strain at peak stress in compression εco′, which
was estimated from the model by Tasemir et al.20 The figure also
shows the concrete strength reduction factor relation proposed
by Collins21 in terms of shear strains, which is essentially the
constitutive model for cracked concrete in compression in the
modified compression field theory (MCFT). This relation has
been shown to relate well to test data in which the compression
stresses act along uniform parallel struts across the section,9 or
elastic shear. Results from this program, however, deviate from
the noted model in interesting ways. The data in Fig. 13(a) shows
how the monotonic and cyclic tests follow different degradation
trends for the softening factor k, with lower values and a faster
decay with increasing shear distortion for the cyclic walls.
The experimental k values are plotted against concrete
strength in Fig. 13(b), where again the data is clearly segre-
gated in terms of the loading protocol. The cyclic and mono-
tonic wall data was fitted with exponential functions only
for the purpose of illustrating the data trends. It can be seen
that the M05 data points are close to each other because
both failed in elastic web crushing. For increasing concrete
strengths, the softening factor decays faster for the cyclic
Fig. 11—Overview of average strain demands on wall webs: walls. It is of interest to note that the softening factor (and
(a) longitudinal strain profiles in M20 walls; and (b) prin- thus, the inelastic web crushing capacity) is affected both
cipal strains versus ductility. by the cyclic loading and the increase in concrete strength,
or, stated differently, cyclic loading had an increased detri-
that increased concrete strength does not lead to additional mental effect on the effective capacity of the shear resisting
deformation capacity beyond a certain point. Wall M20C, struts with increasing concrete compressive strength.
however, failed after completion of two cycles at μΔ = 4, while
M15C failed during the first loading branch of the second Web crushing capacity models
cycle at μΔ = 4. Evaluation of the dissipated hysteretic energy Table 6 compares the web crushing capacities with different
(area inside the hysteresis loop) shows that all walls provided predictive models. It can be noted that ACI shear provisions
essentially the same level of specific energy dissipation for a considerably underestimate the web crushing strength. At the
given displacement ductility level.18 This can be qualitatively same time, the prediction quality of the model by Hines and
seen by observing the hysteretic responses in Fig. 8. Because Seible13 on the cyclic tests deteriorates with increasing concrete
the walls had the same reinforcement details, this is to be strength. Thus, the experimental program revealed that rational
expected. Thus, the increased shear capacity of the HSC walls web crushing models like the one by Hines and Seible need
improved their hysteretic energy dissipation capability. Based further considerations to be applicable to HSC structural walls.
on this evaluation, Wall M20C had 27% higher energy dissi- A modified version of the Hines and Seible model, as well as
pation capacity than Wall M15C.18 The sum of the dissipated the finite element implementation of the MCFT in modeling
energy for all ductility levels shows that increased concrete the behavior of the HSC walls was proposed by Liu,22 and will
strength increased the inelastic energy dissipation capacity of be reported in a future paper by the authors.
the walls by delaying web crushing failure.
Considering the strut resistance mechanism proposed by Hines CONCLUSIONS
and Seible13 (Fig. 1(b)), the similitude in principal strain values Eight cantilever walls were tested with design concrete
(Fig. 11(b)) indicates that the force demand in the walls at equal compressive strengths of 34, 69, 103, and 138 MPa (5, 10,
displacement ductility levels was the same. It is clear, however, 15, and 20 ksi) under cyclic and monotonic loading to study
that the walls had different deformation and web crushing capac- the effects of HSC and damage accumulation on the inelastic
ities. The difference is then attributed to the concrete strength (as web-crushing capacity of structural walls. The following
reflected in Eq. (3)) and the loading pattern. This was explored conclusions are offered specifically referring to structural
by estimating the concrete softening factor k using experimen- walls with well-confined boundary elements:
tally derived values for Lpr and the shear force at web crushing 1. The experiments clearly demonstrated the differences
failure. Equation (1) was used for Walls M05C, M05M, and between elastic and inelastic web crushing behavior and
M10C because they are considered to have failed by elastic web their dependence on concrete compressive strength;

244 ACI Structural Journal/March-April 2014


Fig. 12—Comparison of force-displacement envelopes for cyclic and monotonic tests: (a) M05; (b) M10; (c) M15; and (d) M20.
(Note: 1 kN = 0.225 kip; 1 mm = 0.0394 in.; 1 ksi = 6.895 MPa.)

Fig. 13—Average compression softening factor k versus: (a) normalized maximum shear distortion in plastic region; and
(b) concrete compressive strength.

Table 6—Comparison of web crushing capacities with models


Experiment ACI 31814 Hines and Seible13

Test unit Δu, mm Fu, kN Δu, mm Fu, kN Difference, % Δu, mm Fu, kN Difference, %
M05C 45.0 803 8.64 342 –81 48.5 821 8

M05M 45.0 855 8.13 322 –82 26.9 725 –40

M10C 42.7 751 8.38 387 –80 64.3 751 51

M10M 130 900 9.91 478 –92 101 853 –22

M15C 78.7 819 10.2 497 –87 128 889 62

M15M 133 934 11.7 542 –91 160 966 20

M20C 76.5 815 14.0 589 –82 Flexure

M20M 189 923 13.2 553 –93 196 992 3

Notes: 1 kN = 0.225 kip; 1 mm = 0.0394 in.

2. Increase in concrete compressive strength enhances the While this conclusion may be generally well recognized for
ductility and hysteretic energy capacity of structural walls walls limited by both elastic and inelastic web crushing, exper-
by preventing web crushing shear failures; imental evidence from this research, based on corresponding
3. Web crushing strength increases in proportion to fc′ as long monotonic and cyclic tests, showed that degradation from
as struts remain undamaged. Hence, transformation from an cyclic loading increases with increasing concrete strength;
elastic web crushing failure to an inelastic web crushing failure 5. Concrete compressive strength does not linearly
can be achieved simply by increasing the concrete strength; increase web crushing strength as implied by rational web
4. Damage to struts caused by cyclic loading and inelastic crushing models; rather, the relationship is nonlinear, with a
deformations limits web crushing strength independently of fc′. decreasing limit as concrete strength increases. This obser-

ACI Structural Journal/March-April 2014 245


vation is not well-described by current web crushing models. ACKNOWLEDGMENTS
Therefore, it is advised that these models consider the limits The research described in this paper was carried out under funding from
the National Science Foundation under Grant No. CMS-0530634. The
demonstrated by the reported tests when considering HSC to authors thank the staff and students of MSU’s Civil Infrastructure Labora-
improve the web crushing performance of walls; tory, where the reported work was conducted.
6. The web crushing capacity of structural walls with
well-confined boundary elements was found to be well in REFERENCES
excess of levels acceptable in current practice for a wide 1. Leonhardt, F., and Walther, R., “The Stuttgart Shear Tests 1961,”
Transaction No. 111, Cement and Concrete Association, London, UK, 1961,
range of concrete compressive strengths; and 134 pp.
7. Rational assessment of web crushing limits can open 2. Mattock, A. H., and Kaar, P., “Precast-Prestressed Concrete Bridges,
up new possibilities for acceptable ductile failure modes on 4, Shear Tests of Continuous Girders,” Journal, Portland Cement Associ-
ation Research and Development Labs, V. 3, No. 1, Jan. 1961, pp. 19-46.
reinforced concrete structural walls. 3. Robinson, J. R., “Essais a l’effort trenchant de pouters a ame mince en
béton armé,” Annales des Ponts et Chaussées, Mar.-Apr. 1961, pp. 226-255.
AUTHOR BIOS 4. Placas, A., and Reagan, P. E., “Shear Failure of Reinforced Concrete
ACI member Rigoberto Burgueño is an Associate Professor of structural Beams,” ACI Journal, V. 68, No. 10, Oct. 1971, pp. 763-773.
engineering and Director of the Civil Infrastructure Laboratory at Michigan 5. Wang, T. Y.; Bertero, V. V.; and Popov, E. P., “Hysteretic Behavior
State University, East Lansing, MI. He received his BS, MS, and PhD from of Reinforced Concrete Framed Walls,” Earthquake Engineering Research
the University of California, San Diego, La Jolla, CA. He is a member of ACI Center Report 75/23, University of California, Berkeley, Berkeley, CA,
Committees 341, Earthquake-Resistant Concrete Bridges, and 440, Fiber-Re- Dec. 1975, 367 pp.
inforced Polymer Reinforcement. His research interests include nano-engi- 6. Oesterle, R. G.; Fiorato, A. E.; Johal, L. S.; Carpenter, J. E.; Russell,
neered structural materials, composite materials and structures, multi-scale H. G.; and Corley, W. G., “Earthquake Resistant Structural Walls—Tests
modeling, and seismic performance of reinforced concrete structures. of Isolated Walls,” NSF Report GI-43880, Portland Cement Association,
Skokie, IL, 1976, 315 pp.
ACI member Xuejian Liu is a former Graduate Research Assistant at 7. Oesterle, R. G.; Ariztizabal-Ochoa, J. D.; Fiorato, A. E.; Russell,
Michigan State University, where he received his PhD in civil engineering. H. G.; and Corley, W. G., “Earthquake Resistant Structural Walls—Tests
He is a member of Joint ACI-ASCE Committee 445, Shear and Torsion. of Isolated Walls, Phase II,” NSF Report ENV77-15333, Portland Cement
His research interests include the seismic behavior of reinforced concrete Association, Skokie, IL, 1979, 331 pp.
structures and fiber-reinforced concrete. 8. Vallenas, J. M.; Bertero, V. V.; and Popov, E. P., “Hysteretic
Behavior of Reinforced Concrete Structural Walls,” Earthquake Engi-
ACI member Eric M. Hines is a Principal at LeMessurier Consultants, neering Research Center Report 79/20, University of California, Berkeley,
Inc., Cambridge, MA, and is a Professor of Practice at Tufts University, Berkeley, CA, 1979, 234 pp.
Medford, MA. He received his PhD in structural engineering from the 9. Oesterle, R. G.; Fiorato, A. E.; and Corley, W. G., “Web Crushing
University of California, San Diego. He is a member of ACI Committee 341, of Reinforced Concrete Structural Walls,” ACI Journal, V. 81, No. 3,
Earthquake-Resistant Concrete Bridges. His research interests include the May-June 1984, pp. 231-241.
seismic performance of low-ductility structural systems in moderate seismic 10. Hines, E. M.; Seible, F.; and Priestley, M. J. N., “Seismic Perfor-
regions and inelastic behavior of reinforced concrete structures. mance of Hollow Rectangular Reinforced Concrete Piers with Highly-Con-
fined Corner Elements—Phase I: Flexural Tests, and Phase II: Shear Tests,”
Structural Systems Research Project Report 1999/15, University of Cali-
fornia, San Diego, La Jolla, CA, 1999, 266 pp.
NOTATION 11. Hines, E. M.; Dazio, A.; and Seible, F., “Seismic Performance of
Av = area of transverse steel at given level Hollow Rectangular Reinforced Concrete Piers with Highly-Confined Corner
dθ = incremental angle for calculating critical strut area Elements—Phase III: Web Crushing Tests,” Structural Systems Research Project
Esh = elastic modulus at onset of strain hardening Report 2001/27, University of California, San Diego, La Jolla, CA, 2001, 239 pp.
Fu = shear force at ultimate 12. Hines, E. M.; Restrepo, J. I.; and Seible, F., “Force-Displacement
Fy = ideal yield shear force Characterization of Well Confined Bridge Piers,” ACI Structural Journal,
Fy′ = first yield shear force V. 101, No. 4, July-Aug. 2004, pp. 537-548.
f1 = principal tensile stress 13. Hines, E. M., and Seible, F., “Web Crushing of Hollow Rectangular Bridge
fc′ = unconfined concrete compressive cylinder strength Piers,” ACI Structural Journal, V. 101, No. 4, July-Aug. 2004, pp. 569-579.
fu = ultimate steel stress 14. ACI Committee 318, “Building Code Requirements for Structural
fy = steel yield stress Concrete (ACI 318-11) and Commentary,” American Concrete Institute,
fyv = transverse steel yield stress Farmington Hills, MI, 2011, 503 pp.
jd = distance between flexural tension and compression force resultants 15. ACI Committee 318, “Building Code Requirements for Reinforced
k = concrete compression strength softening factor Concrete (ACI 318-63),” American Concrete Institute, Farmington Hills,
Lpr = plastic hinge region length MI, 1963, 144 pp.
NCfs = flexure-shear strut compression capacity 16. Paulay, T., and Priestley, M. J. N., Seismic Design of Reinforced
NCs = standard shear strut compression capacity Concrete and Masonry Buildings, Wiley Interscience, New York, 1992, 768 pp.
NDfs = flexure-shear strut compression demand 17. Vecchio, F. J., and Collins, M. P., “The Modified Compression-Field
NDs = standard shear strut compression demand Theory for Reinforced Concrete Elements Subjected to Shear,” ACI
R = radius for critical compression strut fan Journal, V. 83, No. 2, Mar.-Apr. 1986, pp. 219-231.
S = transverse reinforcement vertical spacing 18. Liu, X.; Burgueño, R.; Egleston, E.; and Hines, E. M., “Inelastic Web
T = flexural tensile force resultant Crushing Performance Limits of High-Strength-Concrete Structural Wall—
Tyav = effective average flexural tensile yield force resultant Single wall Test Program,” Report No. CEE-RR–2009/03, Michigan State
tw = structural wall thickness University, East Lansing, MI, 2009, 281 pp.
Δu = test unit flexural top displacement 19. Priestley, M. J. N.; Seible, F.; and Calvi, G. M., Seismic Design and
Δy = ideal yield lateral displacement Retrofit of Bridges, John Wiley & Sons, Inc., New York, 1996, 686 pp.
ΔT = incremental tensile flexural force 20. Liu, X., “Inelastic Web Crushing Performance Limits of High-
ε1 = principal tensile strain Strength-Concrete Structural Walls,” PhD dissertation, Department of Civil
εsh = steel strain at onset of strain hardening and Environmental Engineering, Michigan State University, East Lansing,
φy = ideal yield curvature MI, 2010, 215 pp.
μΔ = displacement ductility 21. Tasdemir, M. A.; Tasdemir, C.; Akyuz, S.; Jefferson, A. D.; Lydon,
θfs = flexure-shear crack angle measured from longitudinal axis F. D.; and Barr, B. I. G., “Evaluation of Strains at Peak Stresses in Concrete:
θs = shear crack angle measured from longitudinal axis A Three-Phase Composite Model Approach,” Cement and Concrete
ρh = structural wall transverse reinforcement ratio Composites, V. 20, 1998, pp. 301-318.
ρl = boundary element longitudinal reinforcement ratio 22. Collins, M. P., “Toward a Rational Theory for RC Members in
ρn = structural wall longitudinal reinforcement ratio Shear,” Proceedings, ASCE, V. 104, No. ST4, Apr. 1978, pp. 649-666.
ρs = boundary element transverse reinforcement ratio

246 ACI Structural Journal/March-April 2014


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 111-S21

Response of Precast Prestressed Concrete Circular Tanks


Retaining Heated Liquids
by Michael J. Minehane and Brian D. O’Rourke
The present study investigated the influence of heated water the tank walls will be shown to be prohibitive from a design
storage, upward to 95°C (171°F), on precast prestressed concrete perspective, unless provisions are made for radial displace-
circular tanks. Modern design standards for concrete liquid- ment during service.
retaining structures require that thermal effects be considered
for the serviceability limit state and the ultimate limit state when
RESEARCH SIGNIFICANCE
deemed significant. Most recognized standards, however, do not
This paper investigates the feasibility and implications
provide guidance for the analysis of such effects. Research in this
area is also limited and almost exclusively concerned with ambient of thermal storage using cylindrical concrete reservoirs,
thermal conditions, with a maximum temperature change of 30°C for which there is currently a paucity of information. The
(54°F) in any instance. research has practical applications in the oil, gas, and nuclear
A finite element study incorporating thermomechanical coupling containment industries, in addition to thermal storage for
investigated the magnitude of stresses associated with thermal district heating and related schemes. Although particular
storage. A linear eigenvalue analysis examined the ultimate limit reference is made throughout to precast prestressed concrete
state of buckling for restrained tank walls due to the thermally- storage tanks, the research is also applicable to partially
induced combined axial compression and bending. Consequent prestressed and reinforced concrete reservoirs.
design implications were established and recommendations made
for accommodating thermal loading.
INFLUENCE OF ELEVATED TEMPERATURES ON
Keywords: buckling; elevated temperature; finite element analysis; mate- MATERIAL PROPERTIES
rial properties; prestressed concrete; reservoirs; thermal loading; thermal
storage. Mechanical properties
EN 1992-1-2,5 EN 1992-3,6 and FIB Bulletin 55: Model
INTRODUCTION Code 20107 each define the reductions in the mechanical
The significance of thermal effects on concrete reservoir properties of both the concrete and steel reinforcement for
walls for ambient conditions is long established. An early elevated temperatures. For the temperature range under
study by Priestley1 determined that temperature gradients of consideration for the current study, the associated strength
30°C (54°F) through the wall thickness can exist in warm reductions are insignificant. It is reasonable to suggest that
climates when the effects of solar radiation are considered. any minor reduction in the strength and stiffness of concrete
Priestley1 demonstrated that the resulting tensile stresses may be discounted when the effect of long-term thermal
were large enough to overcome the residual compression, exposure is considered. Mears8 tested concrete specimens
and cracking would inevitably occur. Ghali and Elliott2 subjected to a constant temperature of 65°C (149°F) for
developed closed-form solutions for the thermal analysis 5000 days and observed that the long-term exposure had, in
of elastic tank walls with varying base restraint and that fact, the effect of increasing the compressive strength and
are free at the top. Through numerical examples, it was the modulus of elasticity of concrete. The same trend was
shown that a gradient of 30°C (54°F) through the wall also recorded by Komendant et al.,9 who tested concrete at
thickness was sufficient to cause cracking. This supported 71°C (160°F) for 270 days, and Nasser and Lohtia,10 who
Priestley’s1 proposal that the design should be based on a tested concrete at 121°C (250°F) for 200 days.
serviceability criterion of limiting crack widths rather than It would appear, however, that this trend is only valid for
a limiting tensile stress. Although modern design standards temperatures below 150°C (302°F), as long-term exposure
require that thermal effects are considered for the service- to temperatures in excess of this resulted in a reduction in the
ability limit state, few provide guidance for the analysis of mechanical properties of concrete.10,11
such effects. Pioneering design codes with regard to this
are NZS 31063 and AS 3735,4 which provide design tables, Creep
originally derived by Priestley,1 to calculate hoop forces and Creep of concrete increases at higher temperatures. Exten-
vertical moments for tank walls free at the top and either sive research has been carried out on the influence of
free-sliding, pinned, or fixed at the base. temperature on concrete creep for structures used in nuclear
The studies reviewed were exclusively applicable to tank
ACI Structural Journal, V. 111, No. 2, March-April 2014.
walls free at the top. As thermal storage tanks require a roof, MS No. S-2012-065, doi:10.14359.51686441, was received February 23, 2012, and
the associated radial restraint at the top of the wall alters reviewed under Institute publication policies. Copyright © 2014, American Concrete
Institute. All rights reserved, including the making of copies unless permission is
the internal force distribution. Moreover, the magnitude of obtained from the copyright proprietors. Pertinent discussion including author’s
closure, if any, will be published ten months from this journal’s date if the discussion
the internal forces resulting from the thermal expansion of is received within four months of the paper’s print publication.

ACI Structural Journal/March-April 2014 247


Fig. 1—Creep coefficient multipliers with increasing
Fig. 2—Residual bond strength ratio with increasing
temperature. (Note: °F = 1.8°C + 32.)
temperature. (Note: 1°F = 1.8°C + 32.)
containment. It would appear from the literature that the
Haddad et al.,20 Chang and Tsai,21 Bažant and Kaplan,22 and
use of a thermal scaling factor, or creep coefficient multi-
Huang.23 A wide scatter in the residual bond strength ratios
plier, is appropriate in accounting for temperature effects on
(ratio of bond strength of heated specimen to that of a spec-
creep. Figure 1 presents a comparison of creep coefficient
imen at ambient temperature) is observed. This is primarily
multipliers from guidance provided by CEB 20812 and fib
due to the many variables involved, including concrete
Bulletin 55: Model Code 20107 in addition to experimental
mixture, compressive strength, exposure duration, method
studies carried out by Brown,13 Gross,14 and Nasser and
of cooling, and bar size.
Neville.15 For a temperature of 95°C (203°F), the creep coef-
ficient multipliers range from approximately 1.95 to 2.45.
Stress relaxation
An accurate evaluation of creep at elevated temperatures
Owing to different coefficients of thermal expansion, steel
is difficult to attain, as creep is sensitive to the evaporable
expands relative to concrete with increasing temperature.
water in the mixture. Consequently, an accurate value of
Consequently, for pretensioned members, this will effec-
the moisture content is desirable if a precise assessment is
tively increase the loss of prestress due to stress relaxation.
to be made. The moisture content, particularly at elevated
fib Bulletin 55: Model Code 20107 quantifies the increase
temperatures, is sensitive to the member thickness. The
in loss due to stress relaxation with increasing temperature
majority of the experimental results were developed for the
for a duration of 30 years (Fig. 3). The significance of high
walls of nuclear containment structures, which are generally
temperatures on the stress relaxation is evident, as a value
members of wall thicknesses in the range of 1000 to 1500 mm
of approximately 2.5% at ambient temperature increases to
(39 to 59 in.). Because the walls of precast prestressed
approximately 15.0% at 100°C (212°F).
storage tanks are much thinner—typically 150 to 250 mm
(5.9 to 9.8 in.)—the walls would lose much more moisture
FINITE ELEMENT ANALYSIS
comparably, which would suggest that a lower value of a
Finite element analysis was carried out using commer-
creep coefficient multiplier may be more appropriate. The
cial finite element software. The cylindrical structure was
experimental results are generally based on uniaxial tests.
idealized using two-dimensional (2-D) axisymmetric
As prestressed concrete circular tanks are subject to a multi-
models comprising quadrilateral solid field and continuum
axial state of stress, a further reduction in the predicted
elements. The thermomechanical transient analysis incor-
creep strain would be appropriate, in line with the findings
porated a semicoupled procedure and was time-stepped
of Hannant16 and McDonald.17
according to predefined intervals. A semicoupled analysis
involves running the thermal and structural analyses sepa-
Bond strength
rately and is conducted when the thermal solution is not
Numerous experimental studies conclusively reveal that
considerably affected by changes in geometry. The thermal
the bond strength decreases with increasing temperature.
analysis, which runs first, is governed by the quasi-harmonic
This is primarily attributed to the differing coefficients of
transient heat conduction equation. The resulting tempera-
thermal expansion for steel and concrete. Figure 2 compares
ture distribution for a given time step is subsequently fed to
test results from studies by Harada et al.,18 Kagami et al.,19
the structural analysis for the calculation of displacements

248 ACI Structural Journal/March-April 2014


and consequent stresses and strains. A maximum element Idealization and boundary conditions
size of 0.3 m (0.98 ft) was established following a mesh To idealize a typical precast concrete tank incorporating
convergence study. The material properties used throughout a roof, the wall ends were restrained from radial displace-
the finite element study are given in Table 1. ment but free to rotate. The top of the wall was also free to
displace vertically because the strain due to thermal expan-
Verification of modeling procedure sion far exceeds that imposed by the self-weight of typical
The modeling procedure was verified with existing results roof construction comprising precast flooring on a grid of
in the literature from Priestley,1 Ghali and Elliott,2 and columns. Figure 6 shows the mesh discretization for the
Vitharana and Priestley.24 In each case, the numerical exam- axisymmetric models and displays a deformed contour of
ples considered a tank wall pinned at the base, free at the top, hoop forces obtained from the structural analysis.
and subject to a 30°C (54°F) ambient temperature gradient.
Figure 4 shows the finite element model for the comparison DESIGN IMPLICATIONS
with Priestley1 for a tank of dimensions R = 15.1 m (49.5 ft), Limiting compressive stress
H = 7.2 m (23.6 ft), and t = 0.2 m (7.87 in.). As is evident For a tank restrained radially at its ends, the hoop force
from Fig. 5, good agreement for hoop forces and vertical induced by thermal storage is compressive over the entire
bending moments was observed in each instance. height of wall and is significant in magnitude. It is neces-
sary to check the resulting compressive stress against limits
provided in design standards and guidance. Table 2 pres-
ents compressive stress limits from standards and guidance,
including EN 1992-1-1,25 BS 8007,26 PCI,27 and NZS 3106.3
The limits are expressed in terms of the concrete cylinder
strength fck. The limit stipulated by BS 8007 is given in terms
of the cube strength fcu, but an approximate conversion is
made herein. Although compressive stresses exceeding
those given in Table 2 may not lead to failure, nonlinearity
associated with creep at higher stress-strength ratios would
need to be taken into account.
Table 1—Material properties used in finite element
study
Material property Value
Modulus of elasticity, MPa (psi) 33,000 (4.78 × 106)
Poisson’s ratio 0.2
Coefficient of thermal expan-
10 × 10–6 (5.55 × 106)
sion, /K (/F)
Thermal conductivity, W/mK
1.5 (10.4)
(Btu.in/h.ft2°F)

Fig. 3—Increase in stress relaxation with increasing Specific heat, J/kgK (Btu/lb°F) 900 (0.21)
temperature, after fib Bulletin 55: Model Code 2010.7 (Note: Convective heat transfer coeffi-
8.33 (1.47)
1°F = 1.8°C + 32.) cient, W/m2K (Btu/h ft.2°F)

Fig. 4—Deformed contour of stresses and associated internal force distribution for comparison with Priestley1 (in MPa). (Note:
1 MPa = 145 psi.)

ACI Structural Journal/March-April 2014 249


Figure 7 gives hoop compressive stresses for various presented in this study. As thermal storage tanks commonly
temperatures for a tank with dimensions D = 30.4 m (99.7 ft), incorporate external insulation, the resulting temperature
H = 7.0 m (23 ft), and t = 0.2 m (7.9 in.). A concrete cylinder distribution involves predominantly a constant temperature
strength of 40 MPa (5800 psi) and a modulus of elasticity across the concrete section with a small temperature differ-
of 33 GPa (4744 ksi) is assumed. These section dimensions ential. As such, for simplicity, the small differential that
and material properties are used for each numerical example may be present is ignored herein. Figure 7 shows that an
average temperature of approximately 50°C (90°F) across
Table 2—Limiting compressive stress at service the concrete section produces localized hoop compressive
from modern design standards/guidance stresses at the wall ends that exceed each of the limits given
in Table 2.
Design standard/guidance Limiting compressive stress at service
EN 1992-1-1 0.45fck Circumferential post-tensioning
BS 8007 ≈ 0.41fck (0.33fcu) As the thermally induced hoop forces over the wall
PCI 0.45fck
height are compressive, the circumferential post-tensioning
requirements remain unchanged. The hoop tension arising
NZS 3106 0.40fck from hydrostatic loading when the liquid is not heated is the

Fig. 5—Verification of modeling procedure with Priestley,1 Ghali and Elliott,2 and Vitharana and Priestley.24 (Note: 1 m =
3.28 ft; 1 MPa = 145 psi; 1 kN/m = 5.71 kip/in.; 1 kNm/m = 0.225 kip-ft/ft.)

Fig. 6—Finite element model showing deformed contour of hoop stresses (in MPa). (Note: 1 MPa = 145 psi.)

250 ACI Structural Journal/March-April 2014


Fig. 8—Compressive hoop forces with increasing tempera-
ture. (Note: 1 kN/m = 5.71 kip/in.; 1°F = 1.8°C + 32.)
Fig. 7—Compressive hoop stresses with increasing tempera-
ture. (Note: 1 MPa = 145 psi; 1°F = 1.8°C + 32.)
critical loading condition that the circumferential post-ten-
sioning is designed to cater for. Figure 8 includes the tensile
force distribution resulting from hydrostatic loading derived
using the beam-on-elastic foundation analogy.

Vertical prestressing
The vertical bending moment distribution arising from
hydrostatic and thermally induced loading is given in
Fig. 9. To establish an approximate upper limit on the
vertical moments, a cracking moment was calculated from
the following equation

 I 
M cr = ( ft + fmax )  (1)
 t / 2 

where I is the second moment of area; t is the wall thickness;


ft is the concrete tensile strength; and fmax is the concentric
precompression stress required to eliminate tensile stresses
while also satisfying maximum compressive stress limits at
the extreme fiber.
Figure 10 applies the cracking moment to the previously
derived vertical bending moments. The moments represent
total moments, that is, the hydrostatic moments subtracted
from the thermally induced moments, because the two load-
ings are coexistent. Fig. 9—Vertical bending moments with increasing tempera-
It is apparent that an average temperature across the ture. (Note: 1 kNm/m = 0.225 kip-ft/ft; 1°F = 1.8°C + 32).
concrete section of approximately 40 to 50°C (72 to 90°F)
Influence of creep on thermal response
produces vertical moments that exceed the cracking moment.
The magnitude of strain-induced effects such as thermal,
This temperature limit is approximately consistent with that
creep, and shrinkage stresses at a given time is directly
for the compressive stress requirement.
proportional to the concrete modulus of elasticity at that
time. As such, realistic values of the modulus of elasticity

ACI Structural Journal/March-April 2014 251


Fig. 10—Total vertical moments with increasing temperature. Fig. 11—Influence of creep on concrete modulus of elas-
(Note: 1 kNm/m = 0.225 kip-ft/ft; 1°F = 1.8°C + 32.) ticity in accordance with EN 1992-1-1.25 (Note: 1 GPa =
143.7 ksi.)
are desirable if an accurate evaluation is to be made of the
structural response. A common approach employed by tension-stiffening effects and, depending on the wall thick-
modern design standards involves the use of an age-adjusted ness and reinforcement ratio, can result in load-reduction
modulus of elasticity that includes a creep component. Figure factors exceeding 0.5. An experimental study by Vitharana
11 shows the influence of creep on the concrete modulus et al.28 investigated the moment-curvature response of rein-
of elasticity using the age-adjusted or effective modulus of forced concrete wall elements subject to applied and thermal
elasticity defined in EN 1992-1-125 as follows loading. The study concluded that the ACI 31829 Branson
formulation and the CEB-FIP MC7830 formulation provided
Ecm upper and lower bounds, respectively, for the experimen-
Ec ,eff = (2) tally observed moment-curvature responses. Vitharana et
1 + f ( ∞, to )
al.28 also proposed a modified Branson equation that showed
good agreement with test results for wall elements subject to
Inputs to the EN 1992-1-125 creep model were concrete a simultaneous axial force and flexural moment.
modulus of elasticity = 33 GPa (4744 ksi); wall thickness
t = 200 mm (7.87 in.); relative humidity = 80%; concrete BUCKLING ANALYSIS
compressive cylinder strength = 40 MPa (5800 psi); age It has been established that thermal loading subjects
at loading = 28 days; normal cement class. Figure 11 also restrained tank walls to significant combined axial compres-
includes a thermally adjusted effective modulus of elasticity sion and bending. Because precast prestressed concrete tanks
using a creep coefficient multiplier of 1.5. are essentially shell structures, buckling stability should to
be addressed. For relatively stiff structures, linear eigen-
Influence of cracking on the thermal response value buckling analysis is a technique that can be applied to
Estimating thermal stresses based on elastic section prop- approximate the maximum load that can be sustained prior
erties is only valid for uncracked sections, as is the case to structural instability or collapse. The underlying assump-
for prestressed tanks where decompression is satisfied. tions of a linear eigenvalue buckling analysis are that the
For reinforced concrete or partially prestressed tank walls, linear stiffness matrix remains unchanged prior to buck-
cracking has the effect of considerably relaxing the thermal ling and the stress stiffness matrix is a multiple of its initial
response. NZS 31063 and AS 37354 account for this using value. Accordingly, provided the prebuckling displacements
load-reduction factors that allow for the reduced section have an insignificant influence on the structural response, the
stiffness that accompanies cracking. The factors include technique can be used effectively to predict the load at which
a structure becomes unstable.

252 ACI Structural Journal/March-April 2014


Fig. 12—Buckled shape of cylindrical shell restrained at its ends and subject to thermally induced combined axial compression
and bending.
Commercial finite element software was used to carry out
the linear eigenvalue buckling analysis. The three-dimen-
sional (3-D) models comprised thick shell elements. The
thermally induced compressive hoop forces and vertical
bending moments, derived from the axisymmetric modeling,
were simulated using a combination of internal stress-strain
loading and externally applied radial pressure loading. The
applied loading includes a partial safety factor of 1.55 for
persistent thermal actions in accordance with EN 199031 and
EN 1991-1-5.32 For plate or shell structures, it is prudent to
include an initial geometric imperfection, as the buckling
load is often sensitive to any deviation from the true geom-
etry. Bradshaw33 made efforts to measure concrete cylin-
drical shells in the field and concluded that imperfections
were observed to be as large as the shell thickness. There-
fore, for the current study, an initial geometric imperfection
of the order of magnitude of the shell thickness was adopted
and was represented as out-of-roundness.
The mode of buckling obtained from the finite element
analysis is given in Fig. 12, with the same mode observed for
all tank sizes. The buckled shape displays the characteristic
sinusoidal buckle waves consistent with Koiter’s34 classical
linearized shell buckling theory. The mode shape is sinu-
soidal both axially and circumferentially. Fig. 13—Eigenvalues from finite element linear buckling
Figure 13 presents the eigenvalues extracted from the analysis. (Note: 1°F = 1.8°C + 32.)
finite element study for various average temperatures and
respectively. This arrangement is slightly conservative, as
H2/Dt ratios. The eigenvalues, l, are ratios of the buckling
the use of sufficient external insulation would generally
load to the applied load. An eigenvalue equal to unity indi-
result in a temperature difference between the inside and
cates that structural instability or buckling has occurred.
outside faces of less than 10°C (18°F).
The lowest eigenvalue extracted from the buckling analysis
Comparing the results observed in Fig. 14 with those in
was 2.52. Thus, the ultimate limit state of buckling was not
Fig. 8 and 9, it is evident that the magnitude and significance
reached for the temperature range considered.
of the internal forces are far less for a free-sliding wall. A
noteworthy observation is the hoop tension developed over
FREE-SLIDING CONDITION
the majority of the wall height which, although not exces-
Theoretically, for a free-sliding condition, an average
sive in magnitude, would need to be summed to the hydro-
temperature across the concrete section does not induce
static hoop tension when calculating circumferential post-
any additional stresses. For a gradient experienced across
tensioning requirements.
the wall thickness, however, associated hoop and vertical
bending stresses develop. Figure 14 is an example for a free-
CONCLUSIONS AND RECOMMENDATIONS
sliding wall subject to a temperature distribution resulting
The following conclusions and recommendations may be
from the storage of heated liquids. The inside and outside
drawn from the current study:
temperatures are taken as 95 and 80°C (171 and 144°F),

ACI Structural Journal/March-April 2014 253


Brian D. O’Rourke is a Lecturer and Researcher in the Department of
Civil, Structural and Environmental Engineering at Cork Institute of Tech-
nology. His research interests include structural behavior and materials
technology. He is a Chartered Engineer and member of Engineers Ireland.

NOTATION
D = diameter
Ec,eff = effective or age-adjusted concrete modulus of
elasticity
Ecm = secant concrete modulus of elasticity
fck = concrete compressive cylinder strength
fcu = concrete compressive cube strength
fmax = maximum concentric precompression
stress required to eliminate tensile stresses
ft = concrete tensile strength
H = wall height
I = second moment of area
Mcr = cracking moment
R = radius
T = wall thickness
x = height from base of wall
f(∞,t0) = final creep coefficient
l = eigenvalue

REFERENCES
1. Priestley, M. J., “Ambient Thermal Stresses in Circular Prestressed
Fig. 14—Internal forces resulting from thermal storage Concrete Tanks,” ACI Journal, V. 73, No. 10, Oct. 1976, pp. 553-560.
2. Ghali, E., and Elliott, E., “Serviceability of Circular Prestressed
for a free-sliding wall. (Note: 1 kNm/m = 0.225 kip-ft/ft; Concrete Tanks,” ACI Structural Journal, V. 89, No. 3, May-June 1992,
1 kN/m = 5.71 kip/in.) pp. 345-355.
3. NZS 3106, “Code of Practice for Concrete Structures for Retaining
1. For the most part, the temperature under consideration Liquid,” Standards Association of New Zealand, 2009, 83 pp.
4. AS 3735, “Concrete Structures for Retaining Liquids—Commentary
for the present study does not have a significant adverse (Supplement to AS 3735-2001),” Standards Australia, 2001, 65pp.
effect on the material properties. The most important factors 5. EN 1992-1-2, “General Rules—Structural Fire Design,” Brussels,
that require consideration are creep of the concrete, bond Belgium, 2004.
6. EN 1992-3, “Design of Concrete Structures, Part 3: Liquid Retaining
strength, and stress relaxation for pretensioned and non- and Containment Structures,” Brussels, Belgium, 2006.
pretensioned reinforcement. Where material properties form 7. Fédération Internationale du Béton (fib), “Model Code 2010: Volume
inputs for analysis and design, any associated reduction 1 First Complete Draft,” fib Bulletin 55, Lausanne, Switzerland, 2010,
317 pp.
should be accounted for, particularly if unfavorable. 8. Mears, A. P., “Long Term Tests on the Effect of Moderate Heating
2. For a tank wall restrained radially at its ends, the internal on the Compressive Strength and Dynamic Modulus of Elasticity of
forces resulting from the storage of heated liquids have been Concrete,” Concrete for Nuclear Reactors, SP-34, C. E. Kesler, ed., Amer-
ican Concrete Institute, Farmington Hills, MI, 1972, pp. 355-375.
shown to be significant. As such, a temperature exceeding 9. Komendant, J.; Nicolayeff, V.; Polivka, M.; and Pirtz, D., “Effect
approximately 50°C (90°F) across the concrete section of Temperature, Stress Level, and Age at Loading on Creep of Sealed
appears to be prohibitive based on compressive stress limits Concrete,” Douglas McHenry International Symposium on Concrete and
Concrete Structures, SP-55, B. Bresler, ed., American Concrete Institute,
and vertical prestressing constraints. Consequently, it is Farmington Hills, MI, 1978, pp. 55-81.
recommended that internal insulation be provided to prevent 10. Nasser, K. W., and Lohtia, R. P., “Mass Concrete Properties at High
temperatures from exceeding this. Temperatures,” ACI Journal, V. 68, No. 3, Mar. 1971, pp. 180-186.
11. Carette, G. G., and Malhotra, V. M., “Performance of Dolostone
3. A linear eigenvalue buckling analysis has revealed that and Limestone Concretes at Sustained High Temperatures,” Temperature
the ultimate limit state of buckling for a wall with restrained Effects on Concrete (ASTM STP 858), T. R. Naik, ed., ASTM International,
ends was not reached for the temperature range considered. West Conshohocken, PA, 1985, pp. 38-67.
12. Comité Euro-International du Béton (CEB), “Fire Design of Concrete
A minimum eigenvalue of 2.52 was observed. Structures: in Accordance with CEB/FIP Model Code 90,” Bulletin D’In-
4. Where provisions are made for radial displacements at formation 21, Lausanne, Switzerland, 1991, 120 pp.
the wall ends during service, the 95°C (171°F) maximum 13. Brown, R. D., “Properties of Concrete in Reactor Vessels,” Confer-
ence on Prestressed Concrete Pressure Vessels, Westminster, Mar. 1968,
temperature does not induce excessive stresses. Complica- Paper 13, pp. 131-151.
tions may arise, however, surrounding possible leakage at 14. Gross, H., “High-Temperature Creep of Concrete,” Nuclear Engi-
the joints. Accordingly, it is recommended that a polymer neering and Design, V. 32, No. 1, Apr. 1975, pp. 129-147.
15. Nasser, K. W., and Neville, A. M., “Creep of Concrete at Elevated
liner be included, thereby eliminating concerns regarding Temperatures,” ACI Journal, V. 62, No. 12, Dec. 1965, pp. 1567-1579.
liquid-tightness. 16. Hannant, D. J., “Strain Behaviour of Concrete Up to 95°C under
Compressive Stresses,” Conference on Prestressed Concrete Pressure
Vessels, Westminster, Mar. 1968, pp. 177-192.
AUTHOR BIOS 17. McDonald, J. E., “Creep of Concrete under Various Temperature,
ACI member Michael J. Minehane is a Structural Design Engineer at RPS Moisture and Loading Conditions,” Douglas McHenry International
Group Ltd., Cork, Republic of Ireland. He received his BEng and MEng Symposium on Concrete and Concrete Structures, SP-55, B. Bresler, ed.,
from Cork Institute of Technology, Republic of Ireland, in 2010 and 2011, American Concrete Institute, Farmington Hills, MI, 1978, pp. 31-53.
respectively. His research interests include prestressed concrete, strut- 18. Harada, T.; Takeda, J.; Yamane, S.; and Furumura, F., “Elasticity
and-tie modeling, and finite element analysis. and Thermal Properties of Concrete Subjected to Elevated Tempera-
tures,” Concrete for Nuclear Reactors, SP-34, C. E. Kesler, ed., American
Concrete Institute, Farmington Hills, MI, 1972, pp. 377-406.

254 ACI Structural Journal/March-April 2014


19. Kagami, H.; Okuno, T.; and Yamane, S., “Properties of Concrete 26. BS 8007, “Code of Practice for Design of Concrete Structures for
Exposed to Sustained Elevated Temperatures,” Third International Confer- Retaining Aqueous Liquids,” BSI, London, UK, 1987.
ence on Structural Mechanics in Reactor Technology, Paper H1/5, London, 27. PCI Committee on Precast Prestressed Concrete Storage Tanks,
UK, 1975, pp. 1-10. “Recommended Practice for Precast Prestressed Concrete Storage Tanks,”
20. Haddad, R. J.; Al-Saleh, R. J.; and Al-Akhras, N. M., “Effect of PCI Journal, V. 32, No. 4, 1987, pp. 80-125.
Elevated Temperature on Bond between Steel Reinforcement and Fibre 28. Vitharana, N. D.; Priestley, M. J.; and Dean, J. A., “Behaviour of
Reinforced Concrete,” Fire Safety Journal, V. 43, 2008, pp. 334-343. Reinforced Concrete Reservoir Wall Elements under Applied and Ther-
21. Chiang, C., and Tsai, C., “Time-Temperature Analysis of Bond mally-Induced Loadings,” ACI Structural Journal, V. 95, No. 3, May-June
Strength of a Rebar after Fire Exposure,” Cement and Concrete Research, 1998, pp. 238-248.
V. 33, 2003, pp. 1651-1654. 29. ACI Committee 318, “Building Code Requirements for Reinforced
22. Bažant, Z. P., and Kaplan, M. F., Concrete at High Temperature: Concrete (ACI 318-89) and Commentary,” American Concrete Institute,
Material Properties and Mathematical Models, Longman Group Limited, Farmington Hills, MI, 1989, 353 pp.
England, 1996, 424 pp. 30. Comité Euro-International du Béton/Fédération Internationale du
23. Huang, Z., “Modeling the Bond between Concrete and Reinforcing Béton, “Model Code for Concrete Structures,” third edition, Paris, 1978,
Steel in a Fire,” Engineering Structures, V. 32, 2010, pp. 3660-3669. 348 pp.
24. Vitharana, N. D., and Priestley, M. J., “Significance of 31. EN 1990, “Basis of Structural Design,” Brussels, Belgium, 2005.
Temperature-Induced Loadings on Concrete Cylindrical Reser- 32. EN 1991-1-5, “Actions on Structures—Part 1-5: General Actions—
voir Walls,” ACI Structural Journal, V. 96, No. 5, July-Aug. 1999, Thermal Actions,” Brussels, Belgium, 2008.
pp. 737-749. 33. Bradshaw, R. R., “Some Aspects of Concrete Shell Buckling,” ACI
25. EN 1992-1-1, “Design of Concrete Structures, Part 1: General Rules Journal, V. 60, No. 3, Mar. 1963, pp. 313-328.
and Rules for Buildings,” Brussels, Belgium, 2004. 34. Koiter, W. T., “On the Stability of Elastic Equilibrium,” PhD thesis,
Technological University of Delft, the Netherlands, 1945.

ACI Structural Journal/March-April 2014 255


NOTES:

256 ACI Structural Journal/March-April 2014


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 111-S22

Bond Strength of Spliced Fiber-Reinforced Polymer


Reinforcement
by Ali Cihan Pay, Erdem Canbay, and Robert J. Frosch
To provide increased insight regarding the bond behavior of and incorporated into ACI 440.1R-06 (ACI Committee 440
fiber-reinforced polymer (FRP) bars, 41 glass FRP, carbon FRP, 2006). While 240 specimens were included in the study by
and steel reinforced concrete beams with unconfined tension lap Wambeke and Shield, the majority of the tests had relatively
splices were tested. The test results are analyzed to evaluate the short development lengths, with l/db less than 30, and only
influence of splice length, surface deformation, modulus of elas-
20 splice tests (14 confined splices by Tighiouart et al. [1999]
ticity, axial rigidity, and bar casting position on bond strength.
and six unconfined splices by Mosley et al. [2008]) were
Furthermore, the test results are compared with the current design
expression recommended by ACI Committee 440 to evaluate its included. Splice tests are the preferred test method for the
applicability. This comparison clearly indicates that the current determination of development lengths, but only a few were
design expression is inadequate, and that a new design equation available at the time. Splice tests are preferred because they
is needed. More importantly, however, this research sheds light on provide a “realistic stress-state in the vicinity of the bars”
the importance of the axial rigidity of the reinforcement on bond (ACI Committee 408 2003). It is for this reason that they are
strength. Test results demonstrate that bond strength is linearly used as the basis of the ACI 318-11 (ACI Committee 318
related to the axial rigidity of the reinforcement. This finding 2011) development length expressions.
has future implications regarding the development of improved While knowledge regarding the bond strength of FRP
design expressions and allowing for an improved understanding reinforcement is developing, there is a scarcity of data
of bond strength.
available from splice tests, particularly unconfined splices
Keywords: bond strength; development length; fiber-reinforced polymer that are commonly used in FRP applications. In addition, a
(FRP) reinforcement; reinforced concrete; splice length. number of questions that were highlighted by Mosley et al.
(2008) remain unanswered. In particular, the study recom-
INTRODUCTION mended evaluating the effect of longer splice lengths for
Fiber-reinforced polymer (FRP) reinforcement can FRP reinforcement, as the maximum splice investigated was
provide an alternative solution for structures susceptible to 18 in. (457 mm) for a No. 5 (15.9 mm) bar (l/db = 28.8).
corrosion and where low electric conductivity or magnetic The previous research indicated that bond strength is propor-
transparency is required. FRP bars, however, are anisotropic, tional to the square root of the development length, but it is
have different physical and mechanical properties than that not clear if this trend continues as the development length
of steel reinforcement, and remain linear-elastic until failure. increases. Furthermore, as only No. 5 bars were tested, it is
The modulus of elasticity of glass and aramid FRP bars are not clear if the findings are appropriate for other bar sizes.
approximately one-fifth that of steel. Although carbon FRP This study also found that the surface deformation of the
(CFRP) bars have a higher modulus than glass FRP (GFRP) FRP does not significantly affect bond strength. Different
bars, their modulus is approximately two-thirds that of steel deformations are now available that can shed further light
reinforcing bars. Consequently, design procedures used for on this subject. Finally, the study found that bond strength
steel reinforced members are not necessarily applicable for is related to the modulus of elasticity of the reinforce-
FRP reinforced structures. ment based on tests of steel (E = 29,000 ksi [200 GPa]),
Because the physical and mechanical properties of FRP GFRP (E ≈ 5500 ksi [37.9 GPa]), and AFRP (E ≈ 6800 ksi
bars are different from those of steel reinforcement, espe- [46.8 GPa]). Carbon fiber bars are now available with E
cially the surface deformation and the modulus of elasticity that falls between that of steel and GFRP (E ≈ 20,000 ksi
of the reinforcement, the bond behavior of FRP reinforced [138 GPa]), which may provide additional understanding
concrete specimens is expected to be quite different than regarding this behavior.
that of steel reinforced specimens. An early study on the
first generation of FRP reinforcement by Fish (1992) made RESEARCH SIGNIFICANCE
this fact apparent. The difference in bond strength on The objective of this research study is to provide addi-
modern FRP reinforcement is clearly evident from the tests tional experimental data from splice tests to improve under-
completed by Mosley et al. (2008), which found that glass standing of the bond strength between FRP reinforcement
(GFRP) and aramid (AFRP) bars achieved approximately and concrete. Of particular interest is the influence of a
50% of the stress developed by steel reinforcement for the ACI Structural Journal, V. 111, No. 2, March-April 2014.
same bar size and splice length. Based on these tests, among MS No. S-2012-069.R1, doi:10.14359.51686519, was received June 19, 2012, and
reviewed under Institute publication policies. Copyright © 2014, American Concrete
others, provisions for the development and splices of FRP Institute. All rights reserved, including the making of copies unless permission is
reinforcement were developed (Wambeke and Shield 2006) obtained from the copyright proprietors. Pertinent discussion including author’s
closure, if any, will be published ten months from this journal’s date if the discussion
is received within four months of the paper’s print publication.

ACI Structural Journal/March-April 2014 257


Table 1—Specimen details
LS, in. (mm) L, ft (m) LV, ft (m) LM, ft (m)
12, 18, 24
13.5 (4.11) 3 (0.91) 6 (1.83)
(305, 457, 610)
36 (914) 18 (5.49) 3.75 (1.14) 9 (2.74)
54 (1372) 18 (5.49) 3.75 (1.14) 8.5 (2.59)

for the specimens reinforced with No. 8 (25.4 mm) bars.


Fig. 1—Typical test specimen. Concrete cover and spacing of the reinforcing bars were
maintained constant throughout the experimental program,
and no transverse reinforcement was provided in the constant
moment region. Two No. 3 (9.53 mm) longitudinal steel
bars were provided in the compression side of the beam to
prevent collapse at failure. In addition, the shear spans were
reinforced with No. 3 (9.53 mm) steel stirrups to prevent
shear failure before bond failure.
In addition to the primary variables, several other vari-
ables were considered. While previous tests clearly indi-
cate that the modulus of elasticity is a primary parameter,
it is not clear if this is solely due to the modulus E or due
to the axial rigidity AE of the reinforcement. Therefore, a
specimen was designed where the E of the reinforcement
remains the same using the same reinforcement material, but
the cross-sectional area A is reduced. To maintain the same
Fig. 2—Cross section detail at splice region.
surface area and deformation pattern, a hollow bar was used
number of variables, including splice length, bar size, to reduce A. In addition, top and bottom-cast specimens were
reinforcement modulus of elasticity, and reinforcement constructed to evaluate the influence of the casting position
deformation type. A secondary objective is to evaluate the and allow for linkage between top and bottom-cast results.
applicability of current design expressions for the bond In the top-cast position, the specimens qualify as “top bar”
strength of FRP reinforcement to determine if the current if more than 12 in. (305 mm) of fresh concrete is cast below
expressions can be reliably used for extended splice lengths the reinforcement.
and bar sizes considering the limited range for which the In total, the experimental program consisted of 41 rein-
expressions were developed. forced concrete beam specimens. Details of the specimens
are summarized in Table 2. A four-part notation system is
SPECIMEN DESIGN used to identify the specimens. The specimens were identi-
Five series of beams were tested to evaluate the bond fied first by the descriptive label B (bond) followed by the
strength. The specimens were designed to provide a system- reinforcement type (Tables 3 and 4), the bar size, and finally
atic evaluation of the primary variables and were reinforced by the splice length. For example, B-HC-5-12 stands for a
with either GFRP, CFRP, or steel bars. Both No. 5 and 8 bond test with No. 5 (15.9 mm) CFRP bars spliced at 12 in.
(15.9 and 25.4 mm) bars were also considered, with the (305 mm). Bottom-cast specimens were identified by adding
exception of carbon bars, for which No. 8 bars were not the notation “b” to the splice length.
available. Splice lengths were considered from 12 in.
(305 mm) up to 54 in. (1372 mm). A typical test specimen is MATERIALS
illustrated in Fig. 1, while dimensions for each specimen are
provided in Table 1. Cross-sectional details of the specimens FRP reinforcement
are shown in Fig. 2. FRP reinforcement included glass bars (No. 5 and 8
All beams were rectangular in cross section, with a total [15.9 and 25.4 mm]) and carbon bars (No. 5 [15.9 mm]) with
depth of 16 in. (406 mm). Specimens were designed with a different surface deformations. Three types of bar surfaces
clear spacing of 1 in. (25.4 mm) between the bars located in were considered to evaluate the effect of varying surface
the splice region and with a 1.5 in. (38.1 mm) side and top deformations, including: 1) sand coating; 2) wrapped inden-
clear cover. This limitation represents the minimum clear tations and sand coating; and 3) fabric texture. For glass bars,
spacing and minimum clear cover allowed by ACI 318-11. all three surface types were included. For carbon bars, only
Three reinforcing bars were spliced at the center of the the sand coating and fabric texture surfaces were possible
constant moment region of the beam. The width of the spec- because the wrapped indentation significantly reduces the
imens was controlled by the minimum cover and spacing tensile strength of the bar, which makes this surface treat-
limitations and the size of the reinforcement; therefore, the ment impractical. A general view of the bar surfaces is
width of specimens was 8.75 in. (222 mm) for the specimens shown in Fig. 3. As shown, the same surface type (sand
reinforced with No. 5 (15.9 mm) bars, and 11 in. (279 mm) coated and fabric texture) was obtained for both types of

258 ACI Structural Journal/March-April 2014


Table 2—Specimen details and test results
Series Specimen Bar type Deformation Bar size Ls, in. (mm) Casting position Test age, days Ptest, kip (kN) ftest, ksi (MPa) μavg, psi (MPa)
B-S1-8-18 Steel Deformed No. 8 18 (457) Top 28 33.0 (147) 40.8 (281) 570 (3.93)
B-PG-8-18 Glass Sand-coated No. 8 18 (457) Top 29 24.1 (107) 27.9 (192) 390 (2.69)
Sand and
B-HG-8-18 Glass No. 8 18 (457) Top 31 20.5 (91.2) 23.7 (163) 332 (2.29)
wrapped
Sand and
B-HG1-5-18 Glass No. 5 18 (457) Top 32 14.2 (63.2) 40.7 (281) 357 (2.46)
I wrapped
Sand and
B-HGO-5-18 Glass No. 5 18 (457) Top 35 11.5 (51.2) 32.9 (227) 289 (1.99)
wrapped
B-PG-5-18 Glass Sand-coated No. 5 18 (457) Top 36 16.5 (73.4) 47.3 (326) 415 (2.86)
B-S1-5-18 Steel Deformed No. 5 18 (457) Top 37 24.1 (107) 72.2 (498) 633 (4.36)
B-HC-5-18 Carbon Fabric texture No. 5 18 (457) Top 38 19.9 (88.5) 58.8 (405) 515 (3.55)
B-HC-5-12 Carbon Fabric texture No. 5 12 (305) Top 28 15.1 (67.2) 44.5 (307) 585 (4.03)
B-S1-8-36 Steel Deformed No. 8 36 (914) Top 31 37.1 (165) 57.2 (394) 400 (2.76)
B-PG-8-36 Glass Sand-coated No. 8 36 (914) Top 32 19.9 (88.5) 28.9 (199) 202 (1.39)
Sand and
B-HG-8-36 Glass No. 8 36 (914) Top 34 21.0 (93.4) 30.4 (210) 212 (1.46)
wrapped
II Sand and
B-HG1-5-36 Glass No. 5 36 (914) Top 35 12.4 (55.2) 44.3 (305) 194 (1.34)
wrapped
Sand and
B-HGO-5-36 Glass No. 5 36 (914) Top 38 13.3 (59.2) 47.5 (328) 208 (1.43)
wrapped
B-PG-5-36 Glass Sand-coated No. 5 36 (914) Top 40 13.9 (61.8) 49.9 (344) 219 (1.51)
B-HC-5-36 Carbon Fabric texture No. 5 36 (914) Top 42 22.9 (102) 84.6 (583) 371 (2.56)
Hollow
B-S2-8-12 Steel No. 8 12 (305) Top 133 18.1 (80.5) 29.7 (205) 465 (3.21)
Deformed
B-S1-8-12 Steel Deformed No. 8 12 (305) Top 130 21.9 (97.4) 27.3 (188) 571 (3.94)
B-PG-8-12 Glass Sand-coated No. 8 12 (305) Top 104 17.1 (76.1) 20.0 (138) 418 (2.88)
Sand and
B-HG-8-12 Glass No. 8 12 (305) Top 106 14.0 (62.3) 16.3 (112) 341 (2.35)
III wrapped
B-S1-8-12b Steel Deformed No. 8 12 (305) Bottom 111 21.2 (94.3) 26.3 (181) 551 (3.80)
Sand and
B-HG-8-12b Glass No. 8 12 (305) Bottom 124 14.5 (64.5) 16.9 (117) 354 (2.44)
wrapped
B-PG-8-12b Glass Sand-coated No. 8 12 (305) Bottom 126 15.8 (70.3) 18.4 (127) 385 (2.65)
B-S2-5-24 Steel Deformed No. 5 24 (610) Top 129 23.5 (105) 70.9* (489) 467 (3.22)
B-HC-5-24 Carbon Fabric texture No. 5 24 (610) Top 132 21.9 (97.4) 64.7 (446) 426 (2.94)
B-PC-5-24 Carbon Sand-coated No. 5 24 (610) Top 139 24.1 (107) 71.8 (495) 472 (3.25)
Sand and
B-HG1-5-24 Glass No. 5 24 (610) Top 142 13.6 (60.5) 39.0 (269) 256 (1.77)
wrapped
IV B-HG2-5-24 Glass Fabric texture No. 5 24 (610) Top 148 16.5 (73.4) 47.6 (328) 313 (2.16)
B-PG-5-24 Glass Sand-coated No. 5 24 (610) Top 153 16.7 (74.3) 48.0 (331) 316 (2.18)
Sand and
B-HG1-5-24b Glass No. 5 24 (610) Bottom 155 14.7 (65.4) 42.2 (291) 277 (1.91)
wrapped
B-PG-5-24b Glass Sand-coated No. 5 24 (610) Bottom 161 17.7 (78.7) 50.8 (350) 334 (2.30)
Sand and
B-HG1-5-12 Glass No. 5 12 (305) Top 156 9.5 (42.3) 27.4 (189) 361 (2.49)
wrapped
B-PG-5-12 Glass Sand-coated No. 5 12 (305) Top 160 10.6 (47.2) 30.4 (210) 400 (2.76)
Sand and
B-HG-8-24 Glass No. 8 24 (610) Top 161 20.8 (92.5) 24.1 (166) 253 (1.74)
wrapped
Sand and
B-HG-8-54 Glass No. 8 54 (1372) Top 175 22.8 (101) 33.0 (228) 154 (1.06)
wrapped
Sand and
V B-HG1-5-54 Glass No. 5 54 (1372) Top 177 14.1 (62.7) 50.4 (347) 148 (1.02)
wrapped
B-PG-5-54 Glass Sand-coated No. 5 54 (1372) Top 181 14.1 (62.7) 50.6 (349) 148 (1.02)
B-HC-5-54 Carbon Fabric texture No. 5 54 (1372) Top 183 22.9 (102) 85.0 (586) 249 (1.72)
Sand and
B-HG1-5-12b Glass No. 5 12 (305) Bottom 162 12.1 (53.8) 34.8 (240) 458 (3.16)
wrapped
B-PG-5-12b Glass Sand-coated No. 5 12 (305) Bottom 164 13.6 (60.5) 39.0 (269) 514 (3.54)
Sand and
B-HG-8-24b Glass No. 8 24 (610) Bottom 169 23.0 (102) 26.7 (184) 280 (1.93)
wrapped
*
Reinforcement yielded.

ACI Structural Journal/March-April 2014 259


Table 3—Mechanical properties of fiber-reinforced polymer bars
Bar type Producer Bar size Designation Surface deformation Er, ksi (GPa) fu, ksi (MPa)
HGO Sand and wrapped 5800 (40.0) 71 (490)
No. 5 HG1 Sand and wrapped 6400 (44.1) 98 (676)
Producer 1
HG2 Fabric texture 7300 (50.3) 115 (793)
Glass
No. 8 HG Sand and wrapped 5700 (39.3) 76 (524)
No. 5 PG Sand 6400 (44.1) 89 (614)
Producer 2
No. 8 PG Sand 6200 (42.7) 76 (524)
Producer 1 No. 5 HC Fabric texture 18500 (127.6) 129 (889)
Carbon
Producer 2 No. 5 PC Sand 21700 (149.6) —*
*
Coating of bar peeled at anchor at 100 ksi (689 MPa).

Table 4—Mechanical properties of steel bars


fy, ksi fu, ksi
Bar size Designation Bar type (MPa) (MPa)
S1 Deformed 75 (517) 95 (655)
No. 5
S2 Deformed 60 (414) 100 (689)
S1 Deformed 76 (524) 97 (669)
No. 8 Hollow
S2 76 (524) 97 (669)
deformed

Concrete
The same mixture proportion was used for all test series.
Fig. 3—Carbon, glass FRP, and steel bars. The concrete used a coarse aggregate consisting of river
FRP reinforcement (glass and carbon), allowing for evalua- gravel with a 0.75 in. (19 mm) maximum aggregate size.
tion of the influence of the material type independent of the Batch weights and slump for each series are provided in
surface deformation. Table 5. Concrete compressive and splitting tensile strengths
Tensile tests on representative coupons were performed were obtained from the average of three 6 x 12 in. (152 x
for each type of reinforcement to determine their mechanical 305 mm) cylinders, and are also provided in Table 5.
properties. Coupons for FRP bars were tested considering the Concrete material tests were timed with the testing program
requirements of ACI 440.3R-04 (ACI Committee 440 2004). such that results were obtained on the first, middle, and
The measured modulus of elasticity and ultimate strength of last day of specimen testing for each series. The concrete
the FRP bars are provided in Table 3. Two types of sand and strengths reported are the average over the days of testing
wrapped bars were tested: HGO and HG1. The designation for each test series. As noted in Table 2, which provides the
HGO represents older GFRP bars from the same manufac- concrete age on the day of testing, specimens in each series
turer that were previously tested in the study by Mosley et were tested within a fairly short timeframe. Therefore, the
al. (2008). Over time, the surface deformation on the bars variation of concrete strengths from the average during the
has changed slightly; therefore, two different configurations duration of testing was within 3% for each series. Across all
of the same surface treatment were considered. In addition, series, concrete compressive strength varied from 4010 to
these bars allow for comparison with the earlier study. 5470 psi (27.7 to 37.7 MPa), even though the same concrete
mixture was ordered.
Steel reinforcement
Deformed steel reinforcement consisted of No. 5 and 8 CONSTRUCTION
(15.9 and 25.4 mm) bars, meeting ASTM A615 Grade 60. Specimens in each series were cast at the same time from
To evaluate the effect of axial rigidity, a No. 8 (25.4 mm) the same batch of concrete. The concrete was placed in
hollow reinforcing bar was constructed by drilling a 0.5 in. the forms in two layers, and each layer was vibrated using
(12.7 mm) diameter hole through a 16 in. (406 mm) length mechanical vibrators. The beams were screeded, and the
of the deformed bar to reduce the cross-sectional area in the surface was finished with a magnesium float. The beams
splice region, as shown in Fig. 3. Bars of each size were were covered with wet burlap, and plastic sheets were placed
obtained from the same heat to ensure consistent reinforce- on top of the burlap to prevent moisture loss before final set.
ment material properties in each phase. Table 4 presents the For each series, cylinders were cast simultaneously with the
properties of the steel bars. beams. The cylinders were consolidated, cured, and stored
in the same manner as the test specimens.

260 ACI Structural Journal/March-April 2014


Table 5—Concrete mixture and mechanical properties
Material Series I Series II Series III Series IV Series V
Cement Type I, lb/yd (kg/m )
3 3
425 (252) 430 (255) 429 (255) 430 (255) 428 (254)
Fine aggregate, lb/yd (kg/m )
3 3
1651 (980) 1591 (944) 1611 (956) 1550 (920) 1609 (955)
Coarse aggregate, lb/yd (kg/m )
3 3
1847 (1096) 1849 (1097) 1842 (1093) 1850 (1098) 1842 (1093)
Water, lb/yd3 (kg/m3) 163 (97) 145 (86) 196 (116) 240 (142) 184 (109)
Air, oz/yd3 (mL/m3) 1.1 (43) 1.1 (43) 0 0 0
Water reducer, oz/yd3 (mL/m3) 8.7 (336) 7.8 (302) 6.5 (251) 1.5 (58) 6.4 (247)
Slump, in. (mm) 4 (102) 5 (127) 5.5 (140) 3 (76) 4.5 (114)
fc′, psi (MPa) 5260 (36.3) 5470 (37.7) 4010 (27.7) 4640 (32.0) 4170 (28.8)
ft, psi (MPa) 590 (4.07) 520 (3.58) 380 (2.62) 440 (3.03) 430 (2.96)

Table 6—Influence of casting position


Bar size Splice length Bar type Casting position ftest, ksi (MPa) Ratio, top/bottom
Top 30.4 (210)
PG 0.78
12 in. Bottom 39.0 (269)
(305 mm) Top 27.4 (189)
HG1 0.79
Bottom 34.8 (240)
No. 5
Top 48.0 (331)
PG 0.94
24 in. Bottom 50.8 (350)
(610 mm) Top 39.0 (269)
HG1 0.92
Bottom 42.2 (291)
Top 20.0 (138)
PG 1.09
Bottom 18.4 (127)

12 in. Top 16.3 (112)


HG 0.96
(305 mm) Bottom 16.9 (117)
No. 8
Top 27.3 (188)
S1 1.04
Bottom 26.3 (181)

24 in. Top 24.1 (166)


HG 0.90
(610 mm) Bottom 26.7 (184)
Average 0.93

TEST SETUP AND PROCEDURE bars, and one was attached to the outer reinforcing bar. The
Beams were placed on two supports, and two equal, strains measured with strain gauges and strains calculated
concentrated loads were applied at the end of the cantilever based on flexural theory agreed well; therefore, no strain
with hydraulic rams, creating a constant moment region gauges were installed on the FRP reinforcing bars for the
between the supports as shown in Fig. 1. The rams were remaining series. Strain gauges, however, were installed on
connected to a single hydraulic hand pump to obtain equal all steel reinforced specimens where there was a possibility
pressure in each ram. Load was applied in 0.5 kip (2.22 kN) of yielding the reinforcement.
increments for the specimens with No. 5 (15.9 mm) bars,
and 1 kip (4.45 kN) increments for the specimens with No. 8 STRUCTURAL BEHAVIOR
(25.4 mm) bars. At each load stage, the crack pattern was Specimens with the same beam width and shear span in
mapped, and crack widths were measured on the beam top a given series cracked at approximately the same load. The
surface. Cracks were mapped and measured up to a critical stiffness of the specimens was approximately the same up to
load, beyond which it was considered unsafe to approach the cracking load. Flexural cracks usually first occurred at the
the beam. support or simultaneously at the support and in the constant
Displacements at the ends, supports, and midspan were moment region. As loading increased, further cracks formed
monitored with displacement transducers, while loads were within the constant moment region, shear span, and splice
monitored using load cells. In Series I and II, three strain region. All specimens failed in a brittle side-splitting mode
gauges were placed on the reinforcing bars at the ends of the in the splice region. Two different types of side splitting
splice region. Two were attached to the middle reinforcing were observed during failure. In the first type, the concrete

ACI Structural Journal/March-April 2014 261


Fig. 4—Splitting failure.
cover in the splice region exploded. In the second case, the
failure was not explosive, and the cover remained intact
with the reinforcement. Photographs captured at the time of
failure that illustrate the failure types are shown in Fig. 4.
The splitting plane observed after failure indicated that
failure typically initiated from the splitting cracks present
on the side face. No damage to the surface deformations was
observed on any of the reinforcing bars.
To further illustrate the response, the applied load versus
end deflection curves for Series I are presented in Fig. 5.
Behavior of the specimens can be described by three distinct
stages. In the first stage, before flexural cracking, the
load-deflection curves are linear and the slopes are approx-
imately identical, indicating that before cracking, the stiff-
ness of the specimens is primarily controlled by the concrete. Fig. 5—Load-deflection curves of Series I specimens with
All specimens in a given series cracked at approximately the No. 5 bars.
same load, with slightly higher loads achieved for the stiffer
BOND STRENGTH
bars. In the second stage, after flexural cracking, the slope
The maximum applied load Ptest at the ends of the canti-
reduces; however, the response remains essentially linear up
lever (Fig. 1) and computed reinforcement stress reached at
to failure. In this stage, the flexural stiffness is a function of
failure ftest for each specimen are provided in Table 2. The
the modulus of elasticity and cross-sectional area of the bars.
reinforcement stress at failure was calculated using both
Bars with a lower modulus of elasticity resulted in a lower
cracked section analysis and moment-curvature analysis.
flexural stiffness. In general, beams reinforced with steel
For moment-curvature analysis, the Hognestad stress-strain
bars had the highest stiffness, followed by beams reinforced
curve was used, and the tensile strength of the concrete was
with CFRP bars, and then by GFRP bars. Beams reinforced
neglected. Values from both analyses were approximately
with different types of GFRP bars have approximately the
the same; therefore, stresses from the crack section analysis
same stiffness due to their similar moduli of elasticity. In the
are presented herein. The average bond stress μavg was calcu-
final stage, all specimens failed suddenly by splitting of the
lated assuming that the tension force in the bar is resisted
concrete in the splice region. The same observations were
by a uniform distribution of stress along the surface of the
made for the specimens tested in the other series (Series II
splice. Nominal cross-sectional dimensions were used in all
to V). Load versus deflection plots for all specimens are
calculations. It should be noted that the steel reinforcement
presented in Pay (2005).
in Specimen B-S2-5-24 yielded before splitting failure;
In each series, among the specimens with the same
therefore, its results will not be considered in future analyses.
cross-sectional dimensions and shear span length, steel
The variables investigated in this study are evaluated in
reinforced specimens reached the highest load, followed
the following sections. To eliminate the effect of variations
by CFRP and then GFRP reinforced specimens. Specimens
in the concrete strength, bar stresses and forces are presented
reinforced with GFRP deflected most, followed by CFRP
normalized by the fourth root of the concrete compressive
and steel reinforced specimens. Based on observation, the
strength. As presented in ACI 408R-03 (ACI Committee 408
modulus of elasticity of the reinforcing bar was directly
2003) and Canbay and Frosch (2005), the influence of the
proportional to the failure load, and inversely proportional
compressive strength on bond strength is best represented
to the deflection at the time of failure.
by the fourth root. Considering the differences in concrete
compressive strength in this study, the square root tradi-

262 ACI Structural Journal/March-April 2014


Fig. 6—Effect of splice length on bond strength (No. 5 bars). Fig. 7—Effect of splice length on bond strength (No. 8 bars).

tionally used to represent tensile strength would produce and wrapped bars for No. 5 (15.9 mm) glass bars can be seen
similar findings. in Fig. 6. For this bar size, the results of the sand-coated bars
and the wrapped and sand-coated bars were similar, with
Splice length the sand-coated bars reaching slightly higher bond stresses
The influence of splice length on bond strength was eval- than the wrapped and sand-coated bars except for the 54 in.
uated among specimens with the same bar size and surface (1372 mm) splice specimens, which failed at approximately
type for splice lengths ranging from 12 to 54 in. (305 to the same stress. In addition to these commercially available
1372 mm), and is presented in Fig. 6 for the No. 5 (15.9 mm) reinforcing bars, No. 5 (15.9 mm) fabric texture glass bars
specimens and Fig. 7 for the No. 8 (25.4 mm) specimens. were specifically produced for this test program to evaluate
Best-fit power trend lines are also provided to illustrate the the effect of the bar surface. This bar type was tested using a
trends of the data. As shown, bar stresses reached at failure 24 in. (610 mm) splice (B-HG2-5-24) and reached a normal-
increase as the splice length increases. The effectiveness ized stress of 45.9 ksi (316 MPa), which is essentially the
of increasing the splice, however, decreases as the length same as that achieved with the companion sand-coated bar
increases, as evidenced by the decreasing slope. In addi- (46.3 ksi [319 MPa] for B-PG-5-24). Therefore, GFRP bars
tion, the slope of the curve, which indicates the strength with a fabric surface texture that is considerably smoother
gain provided by increasing the splice length, is different were capable of reaching stresses as high as the sand-coated
for each reinforcement type. The strength gain for the steel GFRP bar. In considering the No. 8 (25.4 mm) bar reinforced
and carbon bars as the splice length increases is significantly specimens (Fig. 7), the same trend is apparent where the
greater than that for the glass bars. For example, doubling sand-coated bars provided similar bond stresses to the sand
the splice length of the No. 5 (15.9 mm) glass bars from and wrapped bars. Slightly higher bond stresses were devel-
18 to 36 in. (457 to 914 mm) increased the stress by only 6%, oped with the sand-coated bars for the shorter splices, with
while the same increase in splice length for the carbon bars approximately the same stress for the longer 36 in. (914 mm)
resulted in a 43% increase in bar stress. In previous studies, splice. Overall, the sand-coated GFRP bars were observed
the influence of splice length for steel (Canbay and Frosch to reach slightly higher stresses for the shorter splice lengths
2005) and short FRP splices (Mosley et al. 2008) was found among the deformation types tested in the experimental
to be proportional to the square root; therefore, increasing program. Considering the minor differences in test results
the splice length from 18 to 36 in. (457 to 914 mm) results in and the variations expected in bond tests, however, the vari-
a 41% increase in bar stress. While the square root is reason- ations in surface deformation produced little difference in
able for the carbon bars, it significantly overestimates the bond strength.
increase for the glass reinforcement, and is not appropriate.
Based on these results, the effect of splice length on the Modulus of elasticity
ultimate stress reached by the reinforcement appears to be The effect of the modulus of elasticity of the reinforce-
a function of the modulus of the elasticity of the reinforce- ment was investigated among the specimens having the
ment. The benefits of an increase in splice length decrease same surface deformation and bar size. Figure 8 shows
as the modulus of elasticity is decreased. As previously the normalized bar stress versus modulus of elasticity
discussed, the bar stress at failure increases as the modulus for No. 5 (15.9 mm) bars with a 24 in. (610 mm) splice.
of elasticity increases. Although two different surface deformations are consid-
ered, the data points follow a linear trend as the modulus of
Surface deformation elasticity increases. Clearly, the modulus of elasticity of the
Three types of surface deformations induced on GFRP reinforcement has a significant influence on the bond strength
bars were tested to evaluate the effect of surface deformation of the reinforcement, with bond strength increasing as the
on splice strength. A comparison of sand-coated versus sand modulus increases.

ACI Structural Journal/March-April 2014 263


Fig. 8—Effect of modulus of elasticity on bond strength. Fig. 9—Effect of axial rigidity on bond strength (No. 5 bars).
(Note: 1 in. = 25.4 mm.) (Note: 1 in. = 25.4 mm.)
Axial rigidity
Axial rigidity of the bar is calculated by multiplying the
nominal cross-sectional area of the bar A and the modulus
of the elasticity of the reinforcement E. The effect of axial
rigidity for No. 5 (15.9 mm) bars is illustrated in Fig. 9,
where the normalized bar force at failure is plotted versus
the axial rigidity AE for the various splice lengths. As shown,
there is an approximately linear trend between axial rigidity
and bar force. The effect of axial rigidity for No. 8 (25.4 mm)
bars is shown in Fig. 10. Carbon bars were not available in
this bar size to enable three points to be plotted across the
horizontal axis; therefore, only two points are available for
the 18 and 36 in. (457 and 914 mm) splice lengths. For the
12 in. (305 mm) splice, however, a specimen (B-S2-8-12) Fig. 10—Effect of axial rigidity on bond strength (No. 8
was constructed that contained the hollow deformed steel bars). (Note: 1 in. = 25.4 mm.)
reinforcing bar in addition to a specimen containing the
identical bar that was not hollowed out (B-S1-8-12). The DeVries et al. (2001) demonstrates that the influence of
bars in both specimens have the same modulus of elasticity, casting position on bond strength is primarily affected by
surface area, and deformation pattern. Due to the reduction in concrete slump and bleeding; therefore, this behavior was
cross-sectional area, however, the axial rigidity was reduced. unexpected. Furthermore, the steel reinforced specimens
As shown in Fig. 10, the hollow steel bar reached a lower bar (No. 8 [25.4 mm] with 12 in. [305 mm] splice) produced
force than that of the solid bar, which indicates the importance essentially the same ratio as the companion FRP reinforced
of axial rigidity on splice strength. Furthermore, the influence specimens. Regardless, in comparing the eight companion
of axial rigidity is shown to again be approximately linear. top and bottom-cast specimens, an average reduction in
Considering these results, the axial rigidity of the reinforce- strength of 7% was observed. The most significant reduction
ment rather than the modulus of elasticity of the reinforce- (22%) was only observed for the No. 5 (15.9 mm) specimens
ment alone is a primary factor influencing splice strength. with a 12 in. (305 mm) splice. As the splice length increased
to 24 in. (610 mm), the ratio decreased resulting in only a
Bar casting position 7% strength reduction. Interestingly, No. 8 (25.4 mm) spec-
Eight bottom-cast specimens were tested along with imens with a 12 in. (305 mm) splice produced essentially no
eight companion top-cast specimens to determine the effect difference in strength. Based on the tests conducted herein,
of casting position on the behavior of specimens with the top-cast bar specimens produced only a minor reduction
lap-spliced reinforcement and provide connection of the in strength. Therefore, the test results from the top-cast spec-
results of the top bar specimens with the large body of test imens can be compared directly with existing test data from
results that exist for bottom-cast lap-splice specimens. Based bottom-cast test specimens. In the worst case, they will only
on the test results (Table 6), the bond strength of top-cast be slightly conservative.
specimens is generally lower than that of the bottom-cast
specimens, as expected. It should be noted, however, that DESIGN EVALUATION
the reinforcement in two of the top-cast specimens reached To evaluate the bond strength of the FRP reinforced
higher bar stresses than the companion bottom-cast speci- specimens, the data was analyzed considering the current
mens even though the slump of the concrete in the series in ACI 440.1R-06 design recommendation where the bar stress
which they were cast (Series III) was the highest. Research ffe can be computed according to Eq. (1). As noted, the term
by Ferguson and Thompson (1962), Jirsa et al. (1982), and C/db should not be taken larger than 3.5. For the specimens

264 ACI Structural Journal/March-April 2014


Fig. 12—Comparison of strength calculations of
Fig. 11—Comparison of strength calculations of CFRP. No. 5 GFRP.

tested in the experimental program conducted herein, C/db


is always less than 3.5, and this limit does not control. The
test results of the FRP specimens were not evaluated with
ACI 318-11 or ACI 408R-03 recommendations because
these expressions were derived for specimens reinforced
with steel deformed bars and, as previously discussed by
Mosley et al. (2008), are not applicable

fc′  l C le  C
f fe =  13.6 e + + 340 where ≤ 3.5 (1)
a  db db db  db

where C is the lesser of the cover to the center of the bar or


one-half the center-to-center spacing of the bars being devel-
oped, in.; le is embedded length of reinforcing bar, in.; and
α is the top bar modification factor (1.5 for reinforcement Fig. 13—Comparison of strength calculations of
placed so that more than 12 in. [305 mm] of fresh concrete No. 8 GFRP.
is cast below the development length or splice; 1.0 for can be expected to be higher than from a direct comparison
other reinforcement). of splice results. In addition, as noted previously, identical
The calculated reinforcement stresses at failure were specimens with the only variable being casting position were
compared with the experimental results. The ratio of the not available. Therefore, other parameters potentially influ-
experimental to calculated stresses for each reinforcement enced the comparison.
type is illustrated in Fig. 11 through 13. The equations were As shown in Fig. 11, for the CFRP reinforced specimens,
evaluated with and without the bar location factor. It should the ratio of the experimental to calculated stress ranges from
be noted that the bar location factor for FRP reinforcement 0.8 to 1.15 when the bar location factor is not considered.
is 1.5 according to ACI 440.1R-06, whereas a factor of 1.3 is With the bar location factor, the equation provides conserva-
currently used by ACI 318-11 and ACI 408R-03. The top bar tive results, and ranges from 1.21 to 1.73. The conservatism
factor of 1.5 was recommended in the work by Wambeke decreases, however, as the splice length increases. While
and Shield (2006) in considering a comparison of eight top ACI 440.1R-06 does not indicate that Eq. (1) does not apply
bar tests relative to the bulk of the data (75 splitting failures for carbon reinforcement, the expression was developed
in total). Unfortunately, no comparison tests were available from the results of only GFRP bars.
to directly evaluate the top bar effect. Based on compari- Although the ACI 440.1R-06 equation was derived from
sons conducted in this research program for specimens with a database of GFRP bars, Fig. 12 and 13 illustrate that the
a depth of 16 in. (406 mm), the maximum top bar effect equation provides unsafe results for the GFRP reinforced
developed was 1.28, with an average of 1.08. Therefore, the specimens in this investigation even with the inclusion
1.5 factor appears overly conservative. This difference in of the bar location factor (α = 1.5). The unconservatism
results may be explained considering that six of the eight increases as the splice length increases from 12 to 54 in.
top bar tests evaluated by Wambeke and Shield (2006) were (305 to 1372 mm). In the case of bottom-cast specimens,
from splice specimens, while the majority of the bottom bar the experimental results are as low as 52% of the calculated
data to which they were compared were from beam end and values for the No. 8 (25.4 mm) bars. Based on comparisons
notched beam tests, which typically produce higher bond of the experimental and calculated results, even with the
strengths. Therefore, the ratio resulting from this analysis

ACI Structural Journal/March-April 2014 265


recommended bar location factor of 1.5, the ACI 440.1R-06 is included. These results clearly indicate that an improved
expression produces significantly unconservative results for design expression for FRP reinforcement is needed.
the GFRP bar reinforced specimens tested in this experi-
mental program. The level of unconservatism varies with the AUTHOR BIOS
splice length, the reinforcement location (if location factor is Ali Cihan Pay is a Structural Engineer at Enka Teknik, Istanbul, Turkey.
He received his BS and MS from Middle East Technical University, Ankara,
included), bar size, and bar type. The ratio of experimental Turkey, and his PhD from Purdue University, West Lafayette, IN.
to calculated results is also observed to generally decrease as
the splice length increases. Erdem Canbay is an Associate Professor of civil engineering at Middle
East Technical University. He received his BS from Istanbul Technical
University, Istanbul, Turkey, and his MS and PhD from Middle East Tech-
CONCLUSIONS nical University.
Based on the results of this research program, the following
Robert J. Frosch, FACI, is a Professor of civil engineering and Associate
conclusions were made: Dean of the College of Engineering at Purdue University. He received his
1. Bond strength is a function of the modulus of elasticity BSE from Tulane University, New Orleans, LA, and his MSE and PhD from
of the reinforcement. As the modulus of elasticity increases, the University of Texas at Austin, Austin, TX. He is a member of the ACI Board
of Direction, is Chair of ACI Subcommittee 318-D, Flexure and Axial Loads
the bond strength linearly increases; (Structural Concrete Building Code), and is a member of ACI Committees
2. Bond strength is a function of the axial rigidity AE of the 224, Cracking; and 318, Structural Concrete Building Code.
reinforcement. As the axial rigidity increases, the bond strength
linearly increases. This relationship suggests that development ACKNOWLEDGMENTS
This study was conducted at both the Karl H. Kettelhut Structural Engi-
of a unified design approach for the development of reinforce- neering Laboratory and Bowen Laboratory for Large Scale Research at
ment, regardless of reinforcement type, is possible; Purdue University, and was made possible under the sponsorship of the
3. Bond strength increases with increasing splice length; Indiana Department of Transportation (INDOT) and the Federal Highway
Administration (FHWA) through the Joint Transportation Research Program
however, this relationship is nonlinear. While previous (JTRP) Project No. SPR-2491. Their support is gratefully acknowledged.
research on relatively short splices (Mosley et al. 2008) Thanks are extended to Hughes Brothers Inc. and Pultrall Inc. for providing
supported a square root relationship between bond strength the FRP reinforcing bars used in this experimental program.
and splice length for FRP reinforcement consistent with that
observed for steel reinforcement (Canbay and Frosch 2005), REFERENCES
ACI Committee 318, 2011, “Building Code Requirements for Structural
the longer splice lengths tested indicate that this relationship Concrete (ACI 318-11) and Commentary,” American Concrete Institute,
is not constant, and depends on the modulus of elasticity of Farmington Hills, MI, 503 pp.
the reinforcement. For GFRP reinforcement, the relationship ACI Committee 408, 2003, “Bond and Development of Straight Reinforcing
Bars in Tension (ACI 408R-03),” American Concrete Institute, Farmington
is significantly lower than the 0.5 power; Hills, MI, 49 pp.
4. Bond strength is essentially independent of the surface ACI Committee 440, 2004, “Guide Test Methods for Fiber-Reinforced
deformation for the FRP reinforcement tested. While three Polymers (FRPs) for Reinforcing or Strengthening Concrete Structures
(ACI 440.3R-04),” American Concrete Institute, Farmington Hills, MI,
significantly different deformation patterns were considered, 40 pp.
similar bond strengths were developed. This finding supports ACI Committee 440, 2006, “Guide for the Design and Construction of
earlier results (Mosley et al. 2008), and also provides support Structural Concrete Reinforced with FRP Bars (ACI 440.1R-06),” Amer-
ican Concrete Institute, Farmington Hills, MI, 44 pp.
that a common design procedure can be used for a variety of ASTM A615/A615M-12, 2012, “Standard Specification for Deformed
FRP bar types. This finding, however, does not imply that and Plain Carbon-Steel Bars for Concrete Reinforcement,” ASTM Interna-
surface deformation is not required. Tests were conducted tional, West Conshohocken, PA, 6 pp.
Canbay, E., and Frosch, R. J., 2005, “Bond Strength of Lap-Spliced
by Pay (2005) using smooth bars. These bars failed at Bars,” ACI Structural Journal, V. 102, No. 4, July-Aug., pp. 605-614.
extremely low stresses in a pullout mode, and illustrate that DeVries, R. A.; Moehle, J. P.; and Hester, W., 1991, “Lap Splice
some level of deformation is required; Strength of Plain and Epoxy-Coated Reinforcements,” Report No. UCB/
SEMM-91/02, University of California, Berkeley, Berkeley, CA, 93 pp.
5. Top bar cast specimens produced only slightly lower Ferguson, P. M., and Thompson, J. N., 1962, “Development Length of
bond strengths (average 7% reduction) than bottom-cast spec- High Strength Reinforcing Bars in Bond,” ACI Journal, V. 59, No. 7, July,
imens. The maximum reduction was 22%, which results in a pp. 887-922.
Fish, K. E., 1992, “Development Length of Fiber-Composite Concrete
maximum top bar factor of 1.28. In addition, similar factors Reinforcement,” master’s thesis, Iowa State University, Ames, IA, 129 pp.
were observed for steel and FRP reinforced specimens, indi- Jirsa, J. O.; Breen, J. E.; Luke, J. J.; and Hamad, B. S., 1982, “Effect of
cating that a different factor is not needed for FRP reinforce- Casting Position on Bond,” International Conference on Bond in Concrete,
Paisley College of Technology, Paisley, Scotland, pp. 300-307.
ment. Based on this research, the 1.5 factor recommended Mosley, C. P.; Tureyen, A. K.; and Frosch, R. J., 2008, “Bond Strength
by ACI Committee 440 (2006) appears overly conservative. of Nonmetallic Reinforcing Bars,” ACI Structural Journal, V. 105, No. 5,
While this research suggests a lower factor, a top bar factor of Sept.-Oct., pp. 634-642.
Pay, A. C., 2005, “Bond Behavior of Unconfined Steel and Fiber
1.3 consistent with ACI 318-11 is recommended; and Reinforced Polymer (FRP) Bar Splices in Concrete Beams,” doctoral
6. The ACI 440.1R-06 equation for the development of dissertation, Purdue University, West Lafayette, IN, 321 pp.
reinforcement results in significantly unconservative results Tighiouart, B.; Benmokrane, B.; and Mukhopadhyaya, P., 1999, “Bond
Strength of Glass FRP Rebar Splices in Beams Under Static Loading,”
for the test results reported herein. For No. 5 (15.9 mm) bars, Construction and Building Materials, V. 13, No. 7, pp. 383-392.
all results are unconservative if the top bar factor (α = 1.5) Wambeke, B., and Shield, C., 2006, “Development Length of Glass
is not included. For the No. 8 (25.4 mm) bars, all results Fiber-Reinforced Polymer Bars in Concrete,” ACI Structural Journal,
V. 103, No. 1, Jan.-Feb., pp. 11-17.
are unconservative regardless of whether the top bar factor

266 ACI Structural Journal/March-April 2014


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 111-S23

Flexural Behavior and Strength of Reinforced Concrete


Beams with Multiple Transverse Openings
by Bengi Aykac, Sabahattin Aykac, Ilker Kalkan, Berk Dundar, and Husnu Can
Reported are the results of experiments on 10 rectangular reinforced to Vierendeel truss action. The deformations in a beam with
concrete (RC) beams with and without multiple web openings. The an opening were shown to increase and the collapse load to
effects of opening geometry, the use of longitudinal stirrups in the decrease as the opening is moved to a more highly stressed
posts between the openings, the use of diagonal reinforcement portion of span. As the opening length and depth increase,
around openings, and the longitudinal reinforcement ratio on the
Mansur et al. (1991) found that the Vierendeel action
flexural behavior of RC beams with openings were investigated.
becomes more pronounced, and the decrease in the collapse
The stirrups in the posts were shown to have a significant contri-
bution to the ductility of an RC beam with openings if no diag- load increases. Mansur et al. (1992) proposed that the deflec-
onal reinforcement is used. For the same reinforcement details, tions of an RC beam with a large rectangular opening can
RC beams with circular openings were found to have higher load be approximately estimated by assigning reduced flexural
capacities and ductilities than beams with rectangular openings. and shear rigidities to the parts of containing the opening.
The experiments indicated that the posts between the openings Tan and Mansur (1996) proposed design guidelines for the
need to be prevented from undergoing shear failure to avoid Vier- strength and serviceability limit states of RC beams with
endeel truss action and allow a beam to develop its ductility and large openings. Mansur (1998) identified different shear
bending capacity. failure modes of RC beams with web openings and devel-
oped design equations. The tests carried out by Tan et al.
Keywords: diagonal reinforcement; plastic mechanism; reinforced concrete
beam; shear failure; shear reinforcement; Vierendeel truss; web crushing (2001) on RC beams with circular openings indicated that
failure; web opening. the use of diagonal reinforcement offers an effective method
in crack control. Mansur (1999) developed design equa-
INTRODUCTION tions for RC beams subject to torsion in addition to bending
Ducts and pipes associated with the mechanical, elec- and shear. The equations correspond to the beam failure as
trical, and sewer systems in a building are usually located a whole, termed as beam-type, and failure of the top and
underneath the floor beams, resulting in a considerable bottom chords separately, termed as frame-type. Mansur et
loss in the usable floor height. Passage of these ducts and al. (2006) concluded that flexural capacities of RC beams
pipes through web openings in floor beams offers an effec- with large circular openings can be closely estimated using
tive way to utilize the entire floor height, providing a more strut-and-tie models. Yang et al. (2006) investigated the
economic and compact design. Nevertheless, the presence strength and behavior of RC deep beams with web open-
of opening(s) in a reinforced concrete (RC) beam reduces ings, and showed that the failure of a deep RC beam is
its load-carrying capacity and increases its service-load caused by the diagonal cracks projecting from the corners
deflections. The studies on concrete beams with transverse of the opening.
openings in the literature focused on providing these beams In all aforementioned studies, RC beams with one or
with strengths and rigidities comparable to their solid coun- two openings were considered. The failure in these beams
terparts by proper reinforcement detailing. In this way, the is generally related to the shear because the openings are
negative effects of the stress concentrations around the usually located in shear spans. In a recent experimental
openings could be eliminated, the load-carrying capacities program (Dundar 2008; Egriboz 2008; Aykac and Yilmaz
increased, and the deflections decreased. 2011), the influence of multiple openings in the span was
Different types of services, including cooling and venti- investigated. The presence of multiple openings was
lation systems, power and sewer systems, and technology assumed to provide a more efficient design by helping the
and communication services, need to be effectively located stress concentrations around openings to be distributed to
and distributed within structures. The presence of multiple the entire beam length. Furthermore, the presence of open-
openings in a beam is needed to accommodate several pipes ings in the central zone in addition to shear spans was
and ducts related to various services. Steel beams with assumed to shift the failure mode of the beam from brittle
multiple web openings (cellular beams) are commonly used shear failure to ductile flexural failure. Attempts were made
for this reason. In this study, the presence of multiple open- to prevent the brittle modes of shear failure (beam-type and
ings in RC beams was considered to improve the design of frame-type), and the ductilities of the beams were increased
RC structures. ACI Structural Journal, V. 111, No. 2, March-April 2014.
In an extensive experimental study on continuous RC MS No. 2012-070.R2, doi:10.14359.51686442, was received October 20, 2012, and
reviewed under Institute publication policies. Copyright © 2014, American Concrete
beams with a large rectangular opening, Mansur et al. (1991) Institute. All rights reserved, including the making of copies unless permission is
established that the failure of these beams is generally related obtained from the copyright proprietors. Pertinent discussion including author’s
closure, if any, will be published ten months from this journal’s date if the discussion
is received within four months of the paper’s print publication.

ACI Structural Journal/March-April 2014 267


Fig. 1—Reinforcement details of specimens without openings. (Note: 1 mm = 0.0394 in.)

Table 1—Test beams EXPERIMENTAL STUDY


Amount of Diagonal
longitudinal Stirrups reinforce- Concrete Test specimens
Beam Opening reinforcement in posts ment batch A total of 10 rectangular RC beams, each 150 mm
RBn No Moderate No No 2 (5.9 in.) wide, 400 mm (15.7 in.) deep, and 4.0 m (13.1 ft)
long, were tested. Four specimens had 200 x 200 mm (7.9 x
RBb No High No No 3
7.9 in.) square openings, and four specimens had Ø200 mm
RRxn Square Moderate No Yes 1 (Ø7.9 in.) circular openings. The reinforcement details of
RRxcn Square Moderate Yes Yes 1 the beams are illustrated in Fig. 1 and 2. In terms of flex-
RRxb Square High No Yes 2 ural reinforcement ratios, the beams denoted with letter “n”
were moderately reinforced (tension reinforcement ratio rt =
RRxcb Square High Yes Yes 2
0.0078), and the beams denoted with letter “b” were heavily
RCb Circular High No No 2 reinforced (rt = 0.014). The letter “x” in the specimen names
RCcb Circular High Yes No 2 corresponds to the presence of Ø10 nonprestressed cables
RCxb Circular High No Yes 1 spiraling around openings (Fig. 3), and “c” corresponds to
short stirrups in posts in longitudinal direction (Table 1).
RCxcb Circular High Yes Yes 1

Material properties
by proper detailing: short stirrups in the chords, and posts Table 2 tabulates the compressive strength of concrete of
and full-depth stirrups next to openings. Furthermore, RC each specimen on the test day obtained from 150 x 300 mm
beams with different opening geometries were tested within (6 x 12 in.) cylinder tests. The mean values and standard
the scope of the program to establish the geometry which deviations of these material tests are tabulated in Table 2
affects the strength and ductility of an RC beam to a lesser together with the number of material tests. The mean values
extent. This paper reports 10 experiments carried out within and standard deviations of the yield and tensile strengths of
the program. The influence of the use of diagonal reinforce- the S420 reinforcing bars and the number of samples for
ment around the openings, the use of stirrups in the posts, each bar size are tabulated in Table 2.
and the opening geometry are the main test parameters. A
comparison of the experimental results with the estimates Test setup and procedure
from the theoretical methods yielded valuable conclusions. A 200 kN (45 kip) capacity steel frame was used for the
tests. The load, applied by a hydraulic cylinder and measured
RESEARCH SIGNIFICANCE by an electronic load cell, was equally distributed to four
This study investigates the effects of different shear loading points by main and secondary spreader beams
reinforcement schemes and the opening geometry on flex- (Fig. 4). In this way, the simply supported beams were
ural behavior of RC beams with multiple transverse open- loaded at two points, each located at a distance of 300 mm
ings. RC beams with different longitudinal reinforcement (11.8 in.) from midspan, and two points, each located at a
ratios were tested, and different failure modes of RC beams distance of 1200 mm (47.2 in.) from midspan. Six-point
with openings were investigated within the course of the bending was adopted instead of four-point bending to more
study. The experimental results were compared with esti- closely simulate the moment distribution in a beam subjected
mates from different theoretical formulations in the literature to uniform distributed loading, which is the most common
to provide background knowledge for establishing design loading condition in real practice. The midspan vertical
rules for RC beams with multiple openings. The findings of deflection, the support settlements, and the distortions in
the present study will also guide further studies in the field.

268 ACI Structural Journal/March-April 2014


Table 2—Results of material tests
Cylinder compressive strength, MPa (ksi) Yield strength, MPa (ksi) Tensile strength, MPa (ksi)
Standard Standard Standard
Material test Mean deviation No. of tests Mean deviation No. of tests Mean deviation No. of tests
RBn 26.8 (3.9) 3.8 (0.6) 2 — — — — — —
RBb 33.9 (4.9) 2.8 (0.4) 4 — — — — — —
RRxn 27.2 (3.9) 4.5 (0.6) 2 — — — — — —
RRxcn 26.3 (3.8) 3.5 (0.5) 2 — — — — — —
RRxb 27.9 (4.0) 1.4 (0.2) 2 — — — — — —
RRxcb 24.8 (3.6) 5.8 (0.8) 2 — — — — — —
RCb 27.8 (4.0) 2.7 (0.4) 2 — — — — — —
RCcb 29.2 (4.2) 2.0 (0.3) 2 — — — — — —
RCxb 28.3 (4.1) 3.6 (0.5) 2 — — — — — —
RCxcb 26.1 (3.8) 6.1 (0.9) 2 — — — — — —
Ø10 bars — — — 476.0 (69.0) 10.2 (1.5) 6 695.7 (100.9) 6.8(1.0) 6
Ø12 bars — — — 550.5 (79.8) 3.6 (0.5) 6 646.0 (93.7) 3.3 (0.5) 6

Fig. 2—Reinforcement details of specimens with openings. (Note: All dimensions in mm; 1 mm = 0.0394 in.)
openings were measured with the help of linear variable
displacement transducers (LVDTs). The load and deflection
measurements were recorded by a data acquisition system.
The beams were loaded up to failure, and the cracks were
marked and the crack widths measured.

FAILURE MODES AND THEORETICAL EQUATIONS

Failure modes of reinforced concrete beams


with openings
Beam- and frame-type shear and web crushing failures
are the three common types of shear failure in RC beams Fig. 3—Diagonal reinforcement spiraling around circular
with openings. In beam-type failure (Fig. 5(a)), a single openings.
crack extending through the entire depth results in failure.
This diagonal crack is assumed to pass through the center crushing failure (Fig. 5(c)) is caused by crushing of concrete
of opening. The frame-type shear failure (Fig. 5(b)) takes between the diagonal cracks.
place when two distinct cracks form in the top and bottom Based on the plastic hinge method, RC beams with open-
chords, and one of the chords fails due to this cracking. Web ings are prone to failure due to formation of a collapse mech-
anism that is composed of four plastic hinges. This type

ACI Structural Journal/March-April 2014 269


Fig. 4—Test setup of present experimental program.
of failure is denoted as Vierendeel truss action (Fig. 5(d))
because the beams behave similar to a Vierendeel panel.
Vierendeel action causes an RC beam to fail at moments
below its bending capacity. Preventing Vierendeel action
and various forms of shear failure ensures that an RC beam
with openings will reach its bending capacity and fail in a
tension-controlled flexural mode due to crushing of concrete
in the compression zone, which is denoted as flexural failure
(Fig. 5(e)).
In the present study, the plastic methods based on truss
analogy (plasticity truss and strut-and-tie methods) were not
used. In these methods, the load capacity of an RC beam
is obtained from the axial capacities of different struts and
ties, obtained from the reinforcement available in each truss
member. RC beams with openings contain numerous B-
and D-regions, and these beams are modeled with a larger
number of truss members compared with RC beams without
openings, causing lengthy and tedious calculations.
Fig. 5—Failure modes of RC beams with multiple
Theoretical equations used in analysis transverse openings.
Shear strength and shear forces in chord members—
Following the ACI approach (ACI Committee 318 2005),  0.092 ⋅ ku ⋅ k p ⋅ ( fc′ + 17.7)  1.61 ⋅ do  
 ⋅ 1 − 
Mansur (1998) was able to develop the following formula  M  h  
Vn = + 0 . 12 ⋅ b ⋅ dv
for beam-type failure  V ⋅d 
 
 +0.846 ρw′ ⋅ f yv 
Vn = 0.17 ⋅ fc′ ⋅ b ⋅ ( d − do ) (2)
Av ⋅ f yv (1)
+ ⋅ ( dv − do ) + Ad ⋅ f yd ⋅ sin a where h is the beam depth; ku is a size coefficient, varying
s from 0.72 to 1.0; kp (Eq. (3)) is a factor accounting for the
reinforcement ratio; M and V are the bending moment and
where fc′ is the concrete strength; b is the beam width; d shear force at critical section, respectively; and rw′ (Eq. (4))
is the effective depth; do is the depth of opening; dv is the is the web reinforcement ratio within dv
distance between the centroids of extreme tension and
0.23
compression reinforcement layers; s is the stirrup spacing;  100 ⋅ As 
k p = 0.82 ⋅  (3)
Av is the area of stirrups; Ad is the cross-sectional area of  b ⋅ d 

the diagonal reinforcement within the failure surface; fyv and
fyd are the yield strengths of the stirrups and diagonal rein-
forcement, respectively; and a is the angle of inclination of
Av + Ad (sin a + cos a )
diagonal reinforcement. ρ′w = (4)
b ⋅ dv
The Architectural Institute of Japan (1988) gives the
following formula for the beam-type shear failure of RC
beams with openings where As is the area of tension reinforcement. Mansur (1998)
modified the maximum allowable shear force (Vu)max formula
of ACI 318 (ACI Committee 318 2005) for RC beams
with openings

270 ACI Structural Journal/March-April 2014


Table 3—Analytical and experimental ultimate load values
Ultimate load, kN (kip)
Beam Failure mode Test Pex Todeschini et al. (1964) Pan ACI Committee 318 (2005) PACI Pex/Pan Pex/PACI Neutral axis depth, mm (in.)
RBn Beam-type flexural 163.8 (36.8) 164.7 (37.0) 160.5 (36.1) 0.99 1.02 59.4 (2.3)
RBb Beam-type flexural 245.8 (55.2) 291.7 (65.6) 289.0 (65.0) 0.84 0.85 95.6 (3.8)
RRxn Vierendeel truss 156.4 (35.2) 159.6 (35.9) 156.9 (35.3) 0.98 1.00 62.5 (2.5)
RRxcn Beam-type flexural 169.3 (38.1) 159.6 (35.9) 156.9 (35.3) 1.06 1.08 62.5 (2.5)
RRxb Vierendeel truss 232.2 (52.2) 288.8 (64.9) 286.8 (64.5) 0.80 0.81 94.2 (3.7)
RRxcb Web crushing 255.1 (57.3) 288.8 (64.9) 286.8 (64.5) 0.88 0.89 94.2 (3.7)
RCb Diagonal tension 269.3 (60.5) 288.8 (64.9) 286.8 (64.5) 0.93 0.94 94.2 (3.7)
RCcb Beam-type flexural 272.3 (61.2) 288.8 (64.9) 286.8 (64.5) 0.94 0.95 94.2 (3.7)
RCxb Beam-type flexural 278.5 (62.6) 289.3 (65.0) 287.2 (64.6) 0.96 0.97 93.3 (3.7)
RCxcb Beam-type flexural 284.1 (63.9) 289.3 (65.0) 287.2 (64.6) 0.98 0.99 93.3 (3.7)

of hinging; and z is the distance between the centroids of


(Vu )max = 5 ⋅ 0.85 ⋅ 0.17 ⋅ fc′ ⋅ b ⋅ ( d − do ) (5)
the chords.

For frame-type shear failure, Mansur (1998) suggested ANALYSIS OF TEST RESULTS
that the shear capacities of both chords should be checked
against the shear forces calculated from the following equa- Failure modes, ultimate loads, ductilities, and
tions, proposed by Nasser et al. (1967) rigidities of beams
Both reference beams (RBn and RBb) underwent
At ( Ab ) tension-controlled flexural failure (Fig. 6). In both beams,
Vut (Vub ) = Vu ⋅ (6) the load was preserved, while the cover concrete crushed
At + Ab
and later dropped suddenly resulting in failure when the top
bars buckled and concrete crushing initiated. In both beams,
where At and Ab are areas of the top and bottom chords, no considerable diagonal cracking took place in shear spans
respectively; Vu is the shear force in the section; and Vut and (Fig. 7(a)). Table 3 indicates that the experimental ultimate
Vub are the shear forces at the top and bottom chords, respec- load of RBn was in close agreement with the values calcu-
tively. Tan and Mansur (1996) suggested that the shear force lated from the rectangular stress block analysis of ACI 318-05
in the section should be distributed to the chords in accor- (ACI Committee 318 2005) PACI and from the Todeschini et
dance to their flexural rigidities rather than their cross-sec- al. (1964) stress-strain model Pan. The load capacity of RBb
tional areas. remained below the bending capacities calculated from both
Flexural modes of failure—The bottom and top chords in models. Furthermore, Table 4 indicates that the ultimate
RC beams with openings are subjected to axial and shear shear forces Vu in both beams at failure were smaller than
forces and bending moments. Due to axial forces and moments their respective shear strength values Vn, implying that shear
in the chords, the liability of an RC beam to develop a failure had no influence on failure.
mechanism composed of four hinges can be evaluated with Table 5 tabulates the deformation ductility index (DDI)
the help of interaction diagrams as established by Tan and and rigidity values of the specimens. DDI is the ratio of a
Mansur (1996) and Mansur and Tan (1999). Considering beam’s deflection at the instant when the applied load drops
that the hinges at the top and bottom chords are subjected to to 85% of the ultimate load to the deflection at yielding of
compression and tension, respectively, and the differences tension reinforcement. DDI is an indicator of the deform-
in the directions of moments at different hinges, an interac- ability of a beam without a significant reduction in load.
tion diagram is prepared for each hinge and checked against The rigidity values in the table correspond to the slope of
the forces and moments that develop in the hinges at service the initial linear branch of the load-deflection curve. In RC
loads. Tan and Mansur (1996) proposed that the axial forces beams, it is quite cumbersome to determine the slope of
in the chords can be obtained from the moment-curvature diagram due to variation of the flex-
ural stiffness along the span caused by the discrete flexural
Mm cracks. Therefore, slope of the load-deflection curve was
Nb = − Nt = (7)
z adapted.

Four different types of failure were observed in beams
with openings. RRxn and RRxb failed due to the forma-
where Nt and Nb are the axial forces in the top and bottom
tion of plastic failure mechanism (Fig. 8(a) and (b)). In
chords, respectively; Mm is the bending moment at the section
both RRxn and RRxb, two hinges formed at the ends of
the top and bottom chords of the opening closest to the end

ACI Structural Journal/March-April 2014 271


distance between the hinges prevented the excessive stress
concentrations around hinges.
Figures 9(b) and (d) illustrate the linear approximations of
the interaction diagrams of RRxn and RRxb, respectively.
Because the top chords are subjected to compression and
bending, the yield planes above the bending moment axis
are composed of two linear segments, which intersect at
the point corresponding to balance failure. The yield planes
below the moment axis correspond to the bottom chords
subjected to tension as well as bending. The points corre-
sponding to the axial forces and moments at the hinges at
failure are also shown in the diagrams. Both figures indi-
cate that all of the points corresponding to the hinges
remain inside the yield surfaces implying that no hinging
was expected at failure. Nevertheless, the plastic hinging
at locations shown in Fig. 9(a) and (c) and failure of RRxn
and RRxb due to hinging might be induced by the excessive
shear deformations in the posts causing additional stresses.
RRxcn, RCcb, RCxb, and RCxcb failed in flexure after
yielding of tension reinforcement (Fig. 8(c)). The final
failure was caused by the crushing of cover and core concrete
and buckling of compression bars. In these beams, the shear
cracks initiated in the chords and posts at the beginning of
loading did not widen and propagate in further stages of
loading, and the flexural cracks at the central part of the
beam controlled the behavior (Fig. 7(a)). Table 3 indicates
that all of these beams failed at loads close to their respec-
tive bending capacities. RCcb, RCxb, and RCxcb exhibited
greater ductilities than their reference beam RBb, with rela-
tive DDI values greater than unity, and had rigidities close to
Fig. 6—Concrete crushing and buckling of compression the rigidity of the reference (Table 5). The neutral axis depth
bars in reference beams. values given in Table 3, calculated using the Todeschini et
al. (1964) stress-strain model, indicate that the compres-
support (Fig. 8(b)). The two beams differed in the loca- sion zone in each beam remained within the top chord up to
tions of the remaining two hinges, which formed at the right failure and was not affected from the openings.
ends of the top and bottom chords of the fourth opening in RCb underwent frame-type shear failure (Fig. 8(d)) due to
RRxn (Fig. 9(a)) and the fifth opening in RRxb (Fig. 9(c)). severe shear cracking in the chords of the opening closest to
Unlike RC beams with one or two openings, the consider- the left end (Fig. 7(a)). Despite the final failure being origi-
able distance between the hinging locations provided that nated from shear, RCb exhibited a ductile flexural behavior
the stresses in the mechanism were distributed to greater up to failure, and the longitudinal reinforcing bars in the
portions of the beam and the reversal of curvature took place beam yielded before failure. This explains why the load
over a longer length. The posts inside the failure mecha- capacity of the beam was only 7% smaller than its calculated
nism were also observed to fail in shear (Fig. 8(b)), which capacity (Table 3), and its DDI value was almost equal to the
was primarily related to the lack of stirrups in the posts. DDI value of the reference beam (Table 5). Table 4 indicates
RRxn failed at an applied load close to its ultimate capacity that the nominal shear strengths of the chords were signifi-
(Table 3), implying that the failure of a moderately reinforced cantly smaller than the shear forces in the chords at failure.
concrete beam with multiple openings due to formation of The use of short stirrups (Fig. 2) in the chords could not
a mechanism does not result in significant reductions in its prevent the frame-type failure. Table 4 shows that RCcb was
load capacity. The formation of mechanism caused greater also liable to frame-type shear failure considering the signif-
reductions in the capacity of RRxb. Table 5 and Fig. 7(b) icant discrepancies between the nominal shear strengths
show that both RRxb and RRxn had ductile behaviors up to of the chords and the shear forces in the chords at failure.
failure, with the DDI value of RRxb only 20% smaller than RCcb, however, failed in flexure, which can be attributed to
RBb, and the DDI of RRxn 40% smaller than RBn. The fact the presence of stirrups in posts.
that RRxn and RRxb exhibited ductilities and load capac- RRxcb failed due to web crushing in the chords above
ities comparable to their respective reference beams origi- and below the second opening (Fig. 7(a) and 8(e)). Due
nated from two reasons. First, the diagonal reinforcement to this failure, RRxcb could not reach its bending capacity
in the beams carried the shear loads in the posts even after (Table 3), and had a limited ductility (Table 5). The failure of
the failure of the posts. Second, the significant longitudinal the beam was not a diagonal tension failure because both the
chords and the beam had adequate shear strengths (Table 4).

272 ACI Structural Journal/March-April 2014


Table 4—Shear force and shear strength values
Beam-type failure Frame-type failure

Vn, kN (kip) Top chord Bottom chord


Beam ACI AIJ Vu, kN (kip) Vnt, kN (kip) Vut, kN (kip) Vnb, kN (kip) Vub, kN (kip)
RBn 168.7 (37.9) — 81.9 (20.6) — — — —
RBb 165.7 (37.2) — 122.9 (27.6) — — — —
RRxn 242.8 (54.6) 141.9 (31.9) 78.2 (17.6) 86.8 (19.5) 39.1 (8.8) 86.8 (19.5) 39.1 (8.8)
RRxcn 242.8 (54.6) 141.9 (31.9) 84.7 (19.0) 86.8 (19.5) 42.3 (9.5) 86.8 (19.5) 42.3 (9.5)
RRxb 257.8 (58.0) 141.8 (31.9) 116.1 (26.1) 94.2 (21.2) 58.1 (13.1) 94.2 (21.2) 58.1 (13.1)
RRxcb 257.8 (58.0) 141.8 (31.9) 127.6 (28.7) 94.2 (21.2) 63.7 (14.3) 94.2 (21.2) 63.7 (14.3)
RCb 147.2 (33.1) 90.0 (20.2) 134.6 (30.2) 38.9 (8.7) 67.3 (15.1) 38.9 (8.7) 67.3 (15.1)
RCcb 147.2 (33.1) 90.0 (20.2) 136.2 (30.6) 38.9 (8.7) 68.1 (15.3) 38.9 (8.7) 68.1 (15.3)
RCxb 257.9 (58.0) 142.1 (31.9) 139.3 (31.3) 94.3 (21.2) 69.6 (15.6) 94.3 (21.2) 69.6 (15.6)
RCxcb 257.9 (58.0) 142.1 (31.9) 142.1 (31.9) 94.3 (21.2) 71.0 (16.0) 94.3 (21.2) 71.0 (16.0)

Fig. 7—(a) Cracking patterns of specimens at failure; and (b) load-deflection curves of specimens.
Nevertheless, both the chords and the entire beam were however, underwent web crushing failure. The maximum
subjected to shear forces above their maximum allowable allowable shear is calculated from Eq. (5) by assuming that
shear forces calculated from Eq. (5), which caused crushing the depth of each chord is constant along its length, which
of concrete between the diagonals. Table 6 indicates that is an over-conservative assumption for circular openings. In
RRxb, RCb, RCcb, RCxb, and RCxcb were also prone to chords above and below circular openings, the chord depth
web crushing because the shear forces at failure exceeded increases from mid-length of the chord to sides. Therefore,
the maximum allowable shear forces. None of these beams, the maximum shear force tolerable by a chord in an RC

ACI Structural Journal/March-April 2014 273


Table 5—Ductilities and rigidities of beams
Deformation ductility
index (DDI) Rigidity
Absolute,
Beam Absolute Relative kN/mm (kip/in.) Relative
RBn 17.8 1.00 10.12 (57.8) 1.00
RRxn 10.7 0.60 5.77 (32.9) 0.57
RRxcn 8.4 0.47 5.87 (33.5) 0.58
RBb 6.4 1.00 12.05 (68.8) 1.00
RRxb 5.6 0.88 8.86 (50.6) 0.74
RRxcb 1.0 0.16 8.62 (49.2) 0.72
RCb 6.1 0.95 8.87 (50.6) 0.74
RCcb 9.5 1.48 9.32 (53.2) 0.77
RCxb 7.8 1.22 10.80 (61.7) 0.90
RCxcb 8.1 1.27 11.03 (63.0) 0.92

ultimate load and rigidity values greater than those of RCcb,


while the DDI value of RCcb exceeded the value of RCxcb.
The rigidity and load capacity of an RC beam with open-
ings can be increased by using diagonal steel if the posts are
reinforced with stirrups, but the diagonal reinforcement does
not contribute to the ductility in this case.

Effect of stirrups in posts on beam behavior


Figure 7(b) illustrates that the stirrups in the posts signifi-
cantly contribute to the energy capacity if no diagonal
reinforcement is used. The DDI value of RCcb (Table 5) is
also considerably greater than the value of RCb, implying
the significant contribution of the stirrups to the ductility in
the absence of diagonal reinforcement. In beams with rect-
angular openings, it appears that the use of stirrups in the
posts does not contribute to the ductility and energy capacity
of the beam if the beam has diagonal reinforcement. The
DDI values of RRxcn and RRxcb were considerably smaller
than the values of RRxn and RRxb, respectively. In beams
with circular openings and diagonal reinforcement, the stir-
rups in the posts have almost no contribution to the ductility.
The DDI value of RCxcb is approximately equal to the value
Fig. 8—Flexural and shear failure modes of specimens with of RCxb. Table 3 indicates that, in all cases, the use of stir-
openings. rups has a positive but minor effect on the load capacity. To
summarize, the stirrups in the posts improve the behavior of a
beam with circular openings is greater than the value from beam when the beam does not have diagonal reinforcement.
Eq. (5). This might be the reason that none of the specimens In the presence of diagonal reinforcement, the stirrups in
with circular openings failed in diagonal compression. Web the posts have little or no contribution to the flexural perfor-
crushing in the chords might have affected the failure of mance. The use of stirrups in the posts in addition to diag-
RRxb, which eventually failed due to Vierendeel action. onal reinforcement causes the posts to be too strong, shifting
the failure to the chords. If the chords fail in brittle modes
Effect of diagonal reinforcement on beam (Vierendeel truss action, diagonal tension, or compression),
behavior the beam exhibits limited ductility.
Figure 10 indicates the load-deflection curves of beams
with and without diagonal reinforcement. The load-deflec- Effect of opening geometry on beam behavior
tion curves indicate that the use of diagonal reinforcement Figure 11 illustrates the load-deflection curves of beams
in an RC beam increases its energy absorption capacity. The with the same reinforcement details but with different
DDI, rigidity (Table 5), and ultimate load (Table 3) values of opening geometries. Both plots indicate that RC beams
RCxb are higher than the values of RCb, implying that diag- with circular openings have much greater energy capacities
onal reinforcement contributes to the flexural behavior if no than the beams with rectangular openings. The DDI and
stirrups are used in the posts. RCxcb, on the other hand, had

274 ACI Structural Journal/March-April 2014


Table 6—Adequacy of beams for shear
Beam-type failure Frame-type failure
(entire section) (chord members)
(Vu)max, kN (Vut)max, kN
Beam (kip) Vu, kN (kip) (kip) Vut, kN (kip)
RBn 248.0 (55.8) 81.9 (20.6) — —
RBb 235.3 (52.9) 122.9 (27.6) — —
RRxn 105.0 (23.6) 78.2 (17.6) 55.2 (12.4) 39.1 (8.8)
RRxcn 105.0 (23.6) 84.7 (19.0) 55.2 (12.4) 42.3 (9.5)
RRxb 106.4 (23.9) 116.1 (26.1) 54.8 (12.3) 58.1 (13.1)
RRxcb 106.4 (23.9) 127.6 (28.7) 54.8 (12.3) 63.7 (14.3)
RCb 106.4 (23.9) 134.6 (30.2) 54.8 (12.3) 67.3 (15.1)
RCcb 106.4 (23.9) 136.2 (30.6) 54.8 (12.3) 68.1 (15.3)
RCxb 107.1 (24.1) 139.3 (31.3) 55.2 (12.4) 69.6 (15.6)
RCxcb 107.1 (24.1) 142.1 (31.9) 55.2 (12.4) 71.0 (16.0)

rigidity values of RCxb and RCxcb are significantly higher


than the values of RRxb and RRxcb, respectively (Table 5).
Furthermore, the ultimate loads carried by RCxb and RCxcb
considerably exceeded the ultimate loads of RRxb and
RRxcb, respectively (Table 3). The less favorable behavior
of RC beams with rectangular openings is mainly due to the
stress concentrations in the corners. In beams with rectan-
gular openings, the shear cracks in the posts were observed
to project from these sharp corners (Fig. 12), and these
cracks caused reductions in the rigidities and load capac-
ities. Secondly, the smaller areas of the circular openings
compared with the rectangular ones resulted in less reduc-
tions in the beam capacities.
Although rectangular openings have more adverse effects Fig. 9—Hinging locations and interaction diagrams of
on the beam behavior compared with circular openings, Beams RRxn and RRxb. (Note: 1 kN = 0.225 kip; 1 kN.m =
provision of rectangular openings in RC beams might be 0.738 kip-ft.)
unavoidable, considering that the air-conditioning ducts in stirrups in the posts increase the ductility and capacity
buildings are usually rectangular. RC beams with rectan- of RC with openings in the absence of diagonal steel.
gular openings should be more carefully designed because • The simultaneous use of diagonal reinforcement and
they are more prone to Vierendeel action and shear failure stirrups in the posts has minor or no contribution to
of the chords. ductilities and load-carrying capacities of RC beams
with openings. The presence of diagonal reinforcement
SUMMARY AND CONCLUSIONS and stirrups in the posts causes the posts to be overly
Two reference beams, four RC beams with multiple strong, which leads to failure of the chords rather than
circular, and four beams with multiple rectangular open- the posts.
ings were tested to determine the flexural performance of • For the same reinforcement details, RC beams with
RC beams with openings. Three of the beams had moderate circular openings have higher ductilities and load
amounts of flexural reinforcement, while the remaining capacities compared with the beams with rectangular
beams were heavily reinforced. Each beam with openings openings. The experiments indicated that the stress
had longitudinal bars and full-depth stirrups adjacent to concentrations at corners of rectangular openings result
openings and short stirrups in the chords. The longitudinal in cracking, which leads to the reductions in the flexural
reinforcement ratio, opening geometry, and use of diagonal rigidities without exhibiting full ductility.
reinforcement and longitudinal stirrups in the posts were the • In RC beams with multiple openings, the consider-
main test parameters. The experiments and comparison with able distance between the hinging locations causes the
theoretical formulations yielded the following conclusions: stresses in the failure mechanism to be distributed to
• The use of diagonal reinforcement contributes to the greater portions of the beam compared with the beams
ductility and load capacity if the posts are not reinforced with one or two openings. Therefore, RC beams with
with stirrups in longitudinal direction. Similarly, the multiple openings exhibit a more ductile behavior
even if they fail due to Vierendeel action. The posts

ACI Structural Journal/March-April 2014 275


Fig. 10—Contribution of diagonal reinforcement to beam
ductility in presence and absence of stirrups in posts. Fig. 11—Contribution of circular opening geometry on beam
ductility.
remaining inside the failure mechanism in Vierendeel
action were observed to fail in shear. Diagonal rein-
forcement around the openings was found to carry the
forces in the posts and prevent the complete failure of
an RC beam, even after the failure of the posts. The
large shear deformations in the posts were shown to
cause Vierendeel action.
• The failure of an RC beam with openings due to Vier-
endeel action causes greater reductions in the load-car-
rying capacity as the longitudinal reinforcement ratio of
the beam increases.
• The use of short stirrups in the chords is not an adequate
measure for prevention of frame-type shear failure. The
use of diagonal reinforcement and stirrups in the posts Fig. 12—Shear cracks projecting from corners of rectan-
limits the extent of shear cracking in the chords, and gular openings.
prevents the frame-type shear failure.
• RC beams with openings are more liable to web ties and ductilities of RC beams with multiple openings can
crushing (diagonal compression) failure if the chords be increased by proper strengthening of the chords and posts.
and beams have high amounts of shear reinforcement When the chords are weak in shear, an RC beam with open-
and small widths. ings is prone to diagonal tension or compression failures
of the chords. The use of short stirrups proved to be effec-
DESIGN RECOMMENDATIONS FOR REINFORCED tive for preventing frame-type shear failure. Furthermore,
CONCRETE BEAMS WITH WEB OPENINGS the chords should be designed to not exceed the maximum
The experimental and analytical results obtained within allowable shear force to prevent web crushing failure. RC
the scope of the present study indicated that the load capaci- beams with rectangular openings are more prone to different
forms of shear failure compared with beams with circular

276 ACI Structural Journal/March-April 2014


openings. Special attention should be given to shear design nology Institute in 1967; his MS from Tulane University, New Orleans, LA,
in 1977; and his PhD from Ankara University, Ankara, Turkey, in 1983.
of beams with rectangular openings. His research interests include behavior of strengthened reinforced concrete
The shear forces are distributed to the chords in accordance to beams under repetitive loading and the behavior of strengthened reinforced
their cross-sectional areas. The most efficient design is achieved concrete columns.
when the openings are placed at mid-depth of the member.
When the centers of openings are offset from mid-depth, one of ACKNOWLEDGMENTS
This paper represents a condensation of the thesis prepared at Gazi
the chords of each opening is weaker than the other, and is more University, Ankara, Turkey, by B. Dundar under the supervision of H. Can
vulnerable to different forms of shear failure. and S. Aykac toward the degree of Master of Science.
The presence of several openings proved to be effective
in improving the behavior of a beam. Unlike RC beams NOTATION
Ab = cross-sectional area of bottom chord
with one or two openings, the stresses inside failure mech- Ad = cross-sectional area of diagonal reinforcement within failure plane
anism were observed to be distributed to greater portions of At = cross-sectional area of top chord
the beam. The posts inside the failure mechanism should Av = cross-sectional area of stirrups
do = diameter of circular opening or depth of rectangular opening
be designed to resist the excessive deformations in the dv = distance between centers of extreme tension and compression
mechanism. The use of longitudinal stirrups in the posts steel layers
or diagonal reinforcement was shown to effectively delay fyd = yield strength of diagonal reinforcement
fyv = yield strength of stirrups
shear failure of the posts and increase the ductility and load z = distance between centroids of top and bottom chords
capacity. The present experiments indicated that the use of a = angle of inclination of diagonal reinforcement with longitudinal
longitudinal stirrups in the posts with a volumetric ratio of beam axis
r t = tension reinforcement ratio
0.016 provided an RC beam with circular openings with a
load capacity and DDI 11 and 48% greater than its reference.
Similarly, the use of diagonal reinforcement with a volu- REFERENCES
ACI Committee 318, 2005, “Building Code Requirements for Structural
metric ratio of 0.017 resulted in an increase of 13 and 22% Concrete (ACI 318-05) and Commentary,” American Concrete Institute,
in the load capacity and DDI of an RC beam with circular Farmington Hills, MI, 430 pp.
Architectural Institute of Japan (AIJ), 1988, “Standard for the Structural
openings compared with its reference. RC beams with Calculation of Reinforced Concrete Structures,” Architectural Institute of
rectangular openings reached load capacities close to their Japan, Tokyo, Japan, 207 pp.
references in the presence of diagonal reinforcement, with a Aykac, S., and Yilmaz, M. C., 2011, “Behaviour and Strength of RC
Beams with Regular Triangular or Circular Web Openings,” Journal of
volumetric ratio of 0.013. Nevertheless, the DDI value of the Faculty of Engineering and Architecture of Gazi University, V. 26, No. 3,
moderately reinforced RC beam with rectangular openings pp. 711-718. (in Turkish)
remained approximately 40% of its reference in this case. Dundar, B., 2008, “Behaviour and Strength of Reinforced Concrete
Beams with Regular Openings,” MSc thesis, Gazi University, Ankara,
It was found that the simultaneous use of longitudinal stir- Turkey, pp. 16-22. (in Turkish)
rups and diagonal steel causes the posts to be overly strong, Egriboz, Y. E., 2008, “Behaviour and Strength of R/C Beams with
causing the chords to fail before the posts. In the present Regular Rectangular or Circular Web Openings,” MSc thesis, Gazi Univer-
sity, Ankara, Turkey, pp. 27-35. (in Turkish)
study, the beams strengthened with both diagonal reinforce- Mansur, M. A., 1998, “Effect of Openings on the Behaviour and Strength
ment and longitudinal stirrups in the posts exhibited ductil- of R/C Beams in Shear,” Cement and Concrete Composites, V. 20, No. 6,
ities more than 50% smaller than the beams with only diag- pp. 477-486.
Mansur, M. A., 1999, “Design of Reinforced Concrete Beams with
onal reinforcement. Small Openings under Combined Loading,” ACI Structural Journal, V. 96,
No. 5, Sept.-Oct., pp. 675-681.
AUTHOR BIOS Mansur, M. A.; Huang, L. M.; Tan, K. H.; and Lee, S. L., 1992, “Deflec-
Bengi Aykac is a Lecturer in the Civil Engineering Department at Gazi tions of Reinforced Concrete Beams with Web Openings,” ACI Structural
University, Ankara, Turkey, where she received her BS, MS, and PhD. Her Journal, V. 89, No. 4, July-Aug., pp. 391-397.
research interests include behavior of concrete beams strengthened by Mansur, M. A.; Lee, Y. F.; Tan, K. H.; and Lee, S. L., 1991, “Tests on
jacketing techniques, concrete beams strengthened with steel plates, and RC Continuous Beams with Openings,” Journal of Structural Engineering,
crack repair by epoxy injection. ASCE, V. 117, No. 6, pp. 1593-1606.
Mansur, M. A., and Tan, K. H., 1999, Concrete Beams with Openings:
Sabahattin Aykac is an Assistant Professor in the Civil Engineering Analysis and Design, CRC Press, Boca Raton, FL, 224 pp.
Department at Gazi University, where he received his BS, MS, and PhD in Mansur, M. A.; Tan, K. H.; and Weng, W., 2006, “Analysis of Concrete
1989, 1993, and 2000, respectively. His research interests include strength- Beams with Circular Web Openings Using Strut-and-Tie Models,” Malay-
ening and repair of concrete beams, earthquake behavior of strengthened sian Journal of Civil Engineering, V. 18, No. 2, pp. 89-98.
concrete beams, and concrete beams with openings. Nasser, K. W.; Acavalos, A.; and Daniel, H. R., 1967, “Behavior and
Design of Large Openings in Reinforced Concrete Beams,” ACI Journal,
Ilker Kalkan is an Assistant Professor in the Department of Civil Engi- V. 64, No. 1, Jan., pp. 25-33.
neering at Kirikkale University, Kirikkale, Turkey. He received his BS from Tan, K. H., and Mansur, M. A., 1996, “Design Procedure for Reinforced
Middle East Technical University, Ankara, Turkey, in 2004, and his MS Concrete Beams with Large Web Openings,” ACI Structural Journal,
and PhD from the Georgia Institute of Technology, Atlanta, GA, in 2006 V. 93, No. 4, July-Aug., pp. 404-411.
and 2009, respectively. His research interests include structural stability, Tan, K. H.; Mansur, M. A.; and Wei, W., 2001, “Design of Reinforced
fiber-reinforced polymer concrete, and strengthening of concrete beams. Concrete Beams with Circular Openings,” ACI Structural Journal, V. 98,
No. 3, May-June, pp. 407-415.
Berk Dundar is a Civil Engineer at Aydiner Construction Company, Todeschini, C. E.; Bianchini, A. C.; and Kesler, C. E., 1964, “Behavior
Turkey. He received his BS from Dokuz Eylul University, Izmir, Turkey, of Concrete Columns Reinforced with High Strength Steels,” ACI Journal,
in 2005, and his MS from Gazi University in 2008. His research interests V. 61, No. 6, June, pp. 701-716.
include the behavior of concrete beams with openings. Yang, K. H.; Eun, H. C.; and Chung, H. S., 2006, “The Influence of Web
Openings on the Structural Behavior of Reinforced High-Strength Concrete
Husnu Can is a Professor in the Civil Engineering Department at Gazi Deep Beams,” Engineering Structures, V. 28, No. 13, pp. 1825-1834.
University, Ankara, Turkey. He received his BS from Ankara Higher Tech-

ACI Structural Journal/March-April 2014 277


NOTES:

278 ACI Structural Journal/March-April 2014


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 111-S24

Experimental Assessment of Inadequately Detailed


Reinforced Concrete Wall Components
by Adane Gebreyohaness, Charles Clifton, John Butterworth, and Jason Ingham
An experimental study undertaken to assess the seismic behavior of pose. Although there have been a number of experimental
reinforced concrete (RC) walls constructed prior to the introduc- studies10-13 where the seismic performance of existing walls
tion of seismic design requirements in the New Zealand Standard was assessed, studies that evaluated the performance of
Model Building By-law is presented. The geometric characteristics lightly and singly reinforced walls with detailing similar
and material properties of the test specimens were replicated from
to those found in the PGC building are limited. Test spec-
those of an existing building. The primary test variables consid-
imens considered in experimental studies reported in the
ered were wall thickness, magnitude of applied axial compressive
load, and aspect ratio. In addition, the influence of the longitu- literature were doubly reinforced,11-13 used higher longitu-
dinal reinforcing bar splices, which are positioned in locations dinal reinforcing bar ratios than those found in many older
that are not permitted by current design standards, on the seismic buildings,10-13 or used deformed bars instead of plain round
performance of the walls was investigated. The response of the test bars.10,11,13 In addition, few of the test specimens investi-
specimens was dominated by rocking, after yielding of the longi- gated previously had spliced longitudinal reinforcement.12
tudinal reinforcing bars located adjacent to the boundaries of the Therefore, it was concluded that a study was necessary to
walls occurred. The peak strength and the stiffness of the test spec- assess the seismic performance of singly and lightly RC
imens dropped rapidly and significantly after low-level drift cycles. walls incorporating inadequate longitudinal bar splices.
Overall, the test specimens exhibited poor ductility and limited The experimental study presented herein was based on the
energy dissipation capacity. Provisions for required tension splice
walls of a case study building located in Wellington, New
lengths of plain round bars in ASCE/SEI 41-06 were found to be
Zealand. This street corner building was constructed in 1928
excessively conservative.
before the publication of NZSS 95,14 which introduced seis-
Keywords: existing buildings; lap splice; lightly reinforced; walls. mic-resistant design requirements in New Zealand for the
first time. The building has internal, one-way, moment-re-
INTRODUCTION sisting, riveted steel frames and RC walls located at the
Reinforced concrete (RC) buildings constructed prior to perimeter as the lateral-force-resisting systems. The walls
the 1970s in New Zealand and other seismically active coun- have limited openings on sides adjacent to neighboring
tries were not designed and detailed to undergo a ductile buildings and have larger and more regular openings on the
mode of failure.1,2 Poor performance of this class of build- street frontages.
ings has been observed in earthquakes occurring in many The case study building was assessed15 in accordance with
countries around the world, such as Chile in 1985,3 Turkey the nonlinear dynamic procedure detailed by ASCE/SEI,2,16
in 1999,4 Chile in 2010,5 and New Zealand in 2011.6 During using a suite of seven earthquake records17 relevant to the
the recent 2010/2011 Canterbury (New Zealand) earth- seismicity of the building site, to determine the performance
quake sequence, this class of buildings performed poorly of the building during a likely earthquake. From the assess-
and in some cases collapsed catastrophically. Christchurch ment, it was found that the capacity of the building will be
City Council Building Safety Evaluation statistics indicate exceeded in moderate level earthquakes, which were defined
that 60% of the pre-1970 RC buildings in Christchurch as being one-third as strong as the design level earthquakes
were deemed suitable for restricted access only or were relevant to the building site.18 The walls were identified to
unsafe after the 6.3 magnitude Christchurch earthquake on be the primary lateral-force-resisting components of the
February 22, 2011, which claimed 181 lives.7 The majority building, and those walls located at the street frontages were
of the fatalities were attributed to the collapse of two RC found to be the most critical walls of the building.
buildings—the PGC building and the CTV building, which To obtain an understanding of the seismic performance
were constructed in 1963 and 1986, respectively.8 The PGC of walls similar to those found in the case study building,
building had RC stair/lift core walls that had similar rein- an experimental program was undertaken on replicas of
forcing bar configurations to the walls discussed herein, and typical segments of the most critical walls. The configu-
the poor performance of the singly and lightly reinforced ration of the reinforcing bars within the walls was deter-
walls led to the collapse of the building.9 mined from the original structural drawings and construc-
Previous research attention has primarily been directed tion specifications of the case study building. The walls are
toward the structural components of pre-1970 RC frame ACI Structural Journal, V. 111, No. 2, March-April 2014.
buildings, principally inadequately detailed columns and MS No. S-2012-071.R2, doi:10.14359.51686520, was received August 13, 2012,
and reviewed under Institute publication policies. Copyright © 2014, American
beam-column joints, owing to the clear seismic risk in Concrete Institute. All rights reserved, including the making of copies unless
this class of buildings that these structural components permission is obtained from the copyright proprietors. Pertinent discussion including
author’s closure, if any, will be published ten months from this journal’s date if the
discussion is received within four months of the paper’s print publication.

ACI Structural Journal/March-April 2014 279


singly reinforced with widely spaced plain round bars, and Table 1—Test matrix
the quantity of longitudinal and transverse reinforcement Test Compressive
within these walls is less than what is required by current specimen lw, mm h, mm t, mm M/Vlw axial load, kN
standards19,20 to induce a ductile response. The longitudinal
WPS1 1300 1750 150 1.35 200
reinforcing bars are spliced just above the floor levels. The
spliced reinforcing bars, which would be required by current WPS2 1300 1750 230 1. 35 300
design standards20 to be hooked, lack proper end anchor- WPS3 1300 2400 150 0.92 0
ages. In addition, the splice lengths are too short according
WPS4 1300 2400 230 0.92 0
to current design standards19,20 to initiate yielding of spliced
reinforcing bars before slip occurs. There is also a lack of WPS5 1300 2400 150 0.92 0
transverse confinement reinforcement to contain concrete in WPS6 1300 2400 230 0.92 0
compression zones and to prevent longitudinal reinforcing WPS7 1300 2400 150 0.92 200
bars from buckling. Similar to the wall piers, the coupling
beams of the case study building, which would typically WPS8 1300 2400 230 0.92 300
receive diagonal reinforcement if constructed to current WPS9 1300 2400 150 0.92 200
standards,21 are reinforced with longitudinal and transverse WPS10 1300 2400 230 0.92 300
reinforcement to resist flexure and shear, respectively.
WSS1 1000 1600 150 0.8 0

RESEARCH SIGNIFICANCE WSS2 1000 1600 230 0.8 0


Because RC buildings constructed before the 1970s Note: 1 mm = 0.0394 in.; 1 kN = 0.225 kip.
comprise a considerable portion of vulnerable RC struc-
tures that pose significant seismic risk, as-built performance Ten of the test specimens (WPS1 to WPS10) were repli-
assessment and seismic retrofitting, preferably without loss cated wall piers and the remaining two test specimens
of heritage attributes to the buildings, are crucial steps toward (WSS1 and WSS2) were replicated coupling beams. Eight
ensuring better performance in future earthquakes. Current of the wall piers (WPS3 to WPS10) and both of the coupling
recommendations2,3,16 for the assessment and improvement beams had boundary reinforcement, as shown in Fig. 1. In
of existing RC walls are based on modern design codes, four of the wall piers (WPS5 to WPS8), the longitudinal
which address new walls and are not necessarily appli- reinforcing bars were spliced near the base of the wall piers,
cable to walls designed and constructed following outdated whereas in the remaining test specimens, the longitudinal
methods. The experimental study presented herein intends reinforcing bars were continuous and were anchored outside
to contribute to a more realistic seismic performance assess- the wall piers.
ment of existing nonconforming RC walls. During construction of the test specimens, a concrete
compressive strength of 21 MPa (3.0 ksi) was speci-
EXPERIMENTAL PROGRAM fied to reflect the average strength determined from core
Twelve full-scale wall components were constructed samples extracted from the case study building. In addition,
in-place and tested in the Civil Test Hall of the University of Grade 300 (fy = 300 MPa [43.5 ksi]) plain reinforcing bars
Auckland as part of a research program planned to assess the were planned to be used, as this grade is close to the strength
seismic performance of existing buildings in New Zealand. of the nominal Grade 240/250 reinforcing bars used during
The test matrix is presented in Table 1. the era of construction of the building. Some of the rein-
forcing bars used, however, were incorrectly supplied as
Description of test specimens Grade 500 (fy = 500 MPa [72.5 ksi]). This error was iden-
The geometric characteristics and reinforcing bar config- tified only after the testing program was completed. Gener-
urations of the test specimens were determined based on ally, Grade 500 bars have a shorter yield plateau, rupture at
the original structural drawings and construction specifi- lower strain levels, require longer development lengths, and
cations of the case study building. The walls in the bottom are more likely to buckle than Grade 300 bars. However,
five stories of the case study building are 230 mm (9 in.) reinforcement grade did not significantly affect the behavior
thick, and the walls in the upper four stories are 150 mm of the test specimens, apart from the apparent influence on
(6 in.) thick. Both types of walls are provided with a single wall strength. Material properties of the test specimens are
layer of reinforcement consisting of 10 mm (3/8 in.) diam- summarized in Table 2.
eter plain round reinforcing bars placed at 305 mm (12 in.)
centers spacing, in both the horizontal and vertical directions Test setup and instrumentation
and located at the midthickness plane of the walls. In addi- The first two wall piers (WPS1 and WPS2) were tested as
tion, two 12 mm (1/2 in.) diameter boundary bars are placed vertical cantilevers (refer to Fig. 2(a)), while the remaining
around all openings. The longitudinal bars are spliced just eight wall piers (WPS3 to WPS10) and the coupling beams
above the floor levels with splice lengths of 305 and 457 mm (WSS1 and WSS2) were subjected to double bending using
(12 and 18 in.) for the 10 and 12 mm (3/8 and 1/2 in.) diam- a steel loading beam mounted on and anchored to the top
eter bars, respectively, with no transverse reinforcing bar RC blocks (refer to Fig. 2(b)). As shown in Fig. 2(c), the
enclosing the lap. In Fig. 1, the dimensions and reinforcing coupling beams were rotated and tested in a vertical orien-
bar arrangements of the test specimens are presented. tation. The double bending loading condition was represen-

280 ACI Structural Journal/March-April 2014


Fig. 1—Test specimen geometries and reinforcement details.

ACI Structural Journal/March-April 2014 281


Table 2—Material properties
Longitudinal reinforcement
Longitudinal and transverse reinforcement Boundary reinforcement splices
Test fc′ Al, At, fy, fyt, fult, lb, φ12,
specimen MPa mm MPa MPa ρl, % ρt, % Ab, mm fy, MPa fult, MPa Splice lb, φ10, mm mm
WPS1 18.4 φ10 530 667 0.20 0.18 — — — No — —
WPS2 20.9 φ10 300 429 0.13 0.12 — — — No — —
WPS3 19.6 φ10 351 488 0.20 0.17 4φ12 388 555 No — —
WPS4 16.2 φ10 351 488 0.13 0.11 4φ12 305 436 No — —
WPS5 29.4 φ10 348 487 0.20 0.17 4φ12 516 662 Yes 305 457
WPS6 24.8 φ10 348 487 0.13 0.11 4φ12 516 662 Yes 305 457
WPS7 21.3 φ10 344 456 0.20 0.17 4φ12 305 438 Yes 305 457
WPS8 22.5 φ10 344 456 0.13 0.11 4φ12 305 438 Yes 305 457
WPS9 20.2 φ10 490 631 0.20 0.17 4φ12 301 433 No — —
WPS10 19.3 φ10 490 631 0.13 0.11 4φ12 301 433 No — —
WSS1 18.7 φ10 351 472 0.21 0.20 4φ12 321 426 No — —
WSS2 21.4 φ10 351 472 0.14 0.13 4φ12 321 426 No — —

tative, as closely as possible, of the fixed-fixed sway support exhibited wide cracks localized at the bottom and at the
condition of wall components in multi-story buildings. top of the walls (refer to Fig. 5) accompanied by rupture
Compressive axial load was applied to the wall piers using of longitudinal reinforcement located adjacent to the wall
four high-strength bars that were positioned parallel to the boundaries. No significant flexural and shear deformations
wall centerline and anchored to the strong floor (refer to were observed or recorded within the body of the walls.
Fig. 2(a) and (b)). The magnitude of the sliding deformations observed at the
Typical test specimen instrumentation is shown schemat- wall-foundation block interfaces was insignificant when
ically in Fig. 3. Load cells, denoted as LC1 to LC3, were compared to the magnitude of wall rocking deformations.
employed to measure the magnitude of forces applied on the No sliding took place at the wall-top block interfaces. There
test specimens. The lateral displacement of the test speci- was visible slipping of the spliced bars along the provided
mens was measured using potentiometer TP1 and the read- splice length, especially during testing of those specimens
ings were corrected to provide the horizontal displacement that were not subjected to axial load. Fracturing of all of the
of the top of the test specimens. Portal gauges PG1-PG6, continuous longitudinal bars was observed during testing,
PG7-PG14, and PG15-PG24 were used to measure rocking, while none of the spliced bars were fractured.
flexural, and shear deformations, respectively. Relative The test specimens that were subjected to double bending
sliding displacements that could have occurred during the responded as intended, with crack development being
tests at the foundation block-specimen, specimen-top block, anti-symmetric about the midheight of the walls at drift
and strong floor foundation block interfaces were also moni- cycles of less than 1%. When the testing progressed to drift
tored using portal gauges PG25, PG26, and PG27, respec- cycles of greater than 1%, the top sections of the test spec-
tively. Ten strain gauges per test specimen were attached to imens underwent significantly larger rotations than did the
the transverse and longitudinal reinforcement. Strain gauges bottom sections, as the top of the test specimens were not
were glued at the midheight of the starter bars to four pairs fully restrained against rotation. This support condition
of spliced bars (one on the starter bar and another one on the resulted in an upward shifting of the inflection point, which
main bar) to monitor the performance of the splices located in turn led to larger bending moments being developed at
within four of the test specimens (WPS5 to WPS8). the base of the walls, causing the cracks at the bottom of the
walls to become wider and the cracks at the top of the walls
Testing procedure to close up. During testing of WPS5 and WPS6, the cracks at
The test specimens were subjected to quasi-static cyclic the top of the walls became fully closed with the longitudinal
loading, with the loading regime based on the ACI-rec- reinforcement that had previously yielded in tension hidden
ommended22 loading sequence for assessing the perfor- in the cracks. Strain gauge readings indicated that the bars
mance of new RC structural components. Potentiometer at those locations had already yielded in tension during early
TP1 (refer to Fig. 3) was employed for the loading regime cycles of loading. At drifts of greater than 1%, damage was
displacement control. localized at the base of the walls and the walls were rocking
on their foundation blocks.
OBSERVED RESPONSE Significant spalling of concrete was observed at all wall
The inelastic response of the test specimens was domi- corners of the thinner test specimens (WPS3, WPS5, WPS7,
nated by rocking (refer to Fig. 4). All of the test specimens and WPS9) (refer to Fig. 5). The thicker test specimens

282 ACI Structural Journal/March-April 2014


Fig. 3—Typical wall instrumentation.

Fig. 2—Test setup.


Fig. 4—Typical components of the lateral displacement of
having axial load (WPS2, WPS8, and WPS10) experienced wall.
relatively limited spalling that was located at the bottom
wall corners only. Spalling of concrete during testing of the RESULTS AND DISCUSSIONS
coupling beam specimens (WSS1 and WSS2) was limited The lateral-force-carrying capacity of the test specimens
to the top wall corners only. The spalling of concrete in was limited by their flexural strength, with the experimen-
the compression zones, which was later exacerbated by tally obtained peak strengths and corresponding drift values
excessive compression at the wall toes as the displacement presented in Table 3. The test specimens with continuous
demand increased, was initiated by buckling of longitu- longitudinal reinforcing bars developed 99 to 108% of their
dinal reinforcement. Buckling of longitudinal reinforcement calculated flexural strength and the test specimens with
was observed during all tests except for those specimens spliced longitudinal reinforcing bars developed 97 to 102%
subjected to no axial load and being 230 mm (9 in.) thick of their computed flexural strength. The calculated flexural
(WPS4 and WPS6). These two specimens also exhibited strengths were determined following routine section analysis
no spalling of concrete (refer to Fig. 5 for the response of procedures and by considering all the reinforcement within
WPS4). Reinforcement grade had no discernible influence the cross section of the test specimens to contribute to the
on reinforcing bar buckling and concrete spalling. strength. The calculations used experimentally determined
reinforcing bar yield strengths and concrete compres-
sive strengths. The strengths of the test specimens with

ACI Structural Journal/March-April 2014 283


Fig. 5—Typical state of test specimens at end of test.
lap-spliced longitudinal reinforcing bars were computed specimens had to maintain their axial-load-carrying capacity
assuming that the splices would develop the yield strength to retain the residual strength to higher drift demands. For
of the spliced bars. example, because of loss of axial-load-carrying capacity,
The strength of the test specimens degraded quickly after WPS9 could not achieve the same level of ductility and
low drift cycles, as shown in Fig. 6, due to rupture of the residual strength as was exhibited by all other wall pier
longitudinal reinforcing bars located adjacent to the wall specimens. Slipping of the spliced bars along the provided
boundaries. For test specimens subjected to applied axial splice length improved the performance of test specimens
compressive load, the lateral-force-carrying capacity of the having spliced longitudinal bars, especially those specimens
test specimens was not completely lost after the longitudinal that were not subjected to compressive axial load (refer to
reinforcing bars had ruptured, due to the flexural strength Fig. 6).
attributable to the applied axial load. However, these test

284 ACI Structural Journal/March-April 2014


Table 3—Measured and calculated strengths of test specimens
Test specimen VTest, kN Drift at VTest, % Vn,SF*, kN Vn,S†, kN Vn,F‡, kN VTest/Vn,SF VTest/Vn,S VTest/Vn,F
WPS1 162 0.96 616 395 151 0.26 0.41 1.07
WPS2 174 0.66 647 513 167 0.27 0.34 1.04
WPS3 159 0.43 404 389 147 0.39 0.41 1.08
WPS4 149 0.20 404 455 148 0.37 0.33 1.01
WPS5 199 0.35 536 424 195 0.37 0.47 1.02
WPS6 194 0.44 536 508 197 0.36 0.38 0.98
WPS7 231 0.36 662 391 233 0.35 0.59 0.99
WPS8 271 0.5 802 492 278 0.34 0.55 0.97
WPS9 260 0.36 740 478 259 0.35 0.54 1.00
WPS10 308 0.21 880 564 310 0.35 0.55 0.99
WSS1 160 0.09 361 320 149 0.44 0.50 1.07
WSS2 150 0.04 361 417 151 0.42 0.36 0.99
Average 0.36 0.45 1.02
Standard deviation 0.05 0.09 0.04
*
Nominal shear friction capacity according to ACI 318-11.

Nominal shear capacity according to ACI 318-11.

Nominal flexural capacity according to ACI 318-11 and assuming splices would develop yield strength of spliced bars.
Note: 1 kN = 0.225 kip.

The boundary and distributed longitudinal reinforcement Flexural failure is the principal failure mode for existing
provided to the walls were below the limits that are specified RC walls constructed in New Zealand before the amend-
by both ACI 318 and NZS 3101 (refer to Table 2). ACI 318 ment of NZSS 9514 in 1955, mainly because in NZSS 95
stipulates the minimum longitudinal reinforcement ratio the contribution of concrete to the shear strength of RC
for earthquake force-resisting RC walls to be 0.25% if Vu structural components is underestimated.23 Consequently, in
exceeds 0.83Acv√fc′ (MPa) (10Acv√fc′ [psi]), and the tested the absence of boundary frame elements, walls of this era
walls satisfy this condition. In addition to the minimum are generally expected to have sufficient shear strength to
ratio, ACI 318 requires at least two 16 mm (5/8 in.) diameter develop flexural overstrength. However, due to the provision
bars to be provided around all openings. The requirement of a low quantity of reinforcement and a lack of transverse
of NZS 3101 is dependent on concrete strength and rein- confinement reinforcement, this type of wall has low flexural
forcement yield strength. For the walls tested, the minimum strength and exhibits a low ductility capacity.
required reinforcement ratio was, on average, 0.3%. The test specimens dissipated a significant amount of
NZS 3101 also requires additional reinforcement with yield energy through yielding of longitudinal reinforcing bars
strength equal to or greater than 600 N/mm (3426 lb/in.) of wall during cycles to drifts of less than 1%. However, the
thickness to be provided around all openings. This require- energy dissipation capacity of the test specimens reduced
ment is approximately equivalent to two 12 mm (1/2 in.) considerably as the drift demands increased, due to a lack
and three 12 mm (1/2 in.) diameter Grade 500 bars for the of any dissipative mechanism except sliding friction at the
150 and 230 mm (6 and 9 in.) thick walls, respectively. wall-foundation block interfaces. The bar slip that occurred
Due to the additional two 12 mm (1/2 in.) diameter over the splice lengths did not significantly alter the energy
boundary reinforcing bars, the wall pier specimens that were dissipation capacity for those test specimens that incorpo-
tested in double bending (WPS3 to WPS10) achieved more rated lap splices.
strength and energy dissipation when compared to the wall After 1% drift demand, the stiffness of the test speci-
pier specimens having no boundary reinforcement (WPS1 mens had typically degraded significantly to less than 10%
and WPS2). However, even with these additional bars, the of the initial stiffnesses, with continuing degradation until
quantity of longitudinal reinforcement provided resulted in reaching approximately 1% of the initial stiffnesses at drift
a level of applied shear force that was insufficient to induce demands in excess of 2.5%. The rapid and significant loss of
a shear mode of failure. Similar findings were previously stiffness observed during testing indicates that, after large
reported in studies12,13 conducted on the behavior of existing earthquakes, this type of wall becomes too soft to develop
RC walls detailed following pre-1970s’ detailing techniques, significant ongoing resistance and therefore will undergo
but having relatively more longitudinal reinforcement than larger displacements when subjected to small lateral forces
used in the test specimens discussed herein. associated with either a long-duration event or aftershocks
having significant intensity at the site.

ACI Structural Journal/March-April 2014 285


Fig. 6—Lateral-force, top-displacement responses of test specimens.
COMPARISON OF RESULTS WITH ASCE 41-06 aspect ratios ranging between 1.35 and 1.85, the response
PROVISIONS of the walls was supposed to be controlled by shear and by
Section C6.7.1 of ASCE/SEI 41-06,16 “Seismic Rehabili- both shear and flexure, respectively, according to ASCE/SEI
tation of Existing Buildings,” categorizes the response of RC 41-06.
walls based on aspect ratio. According to ASCE/SEI 41-06,
walls with an aspect ratio of less than 1.5 are considered to Wall strength
be squat and their response is assumed to be controlled by ASCE/SEI 41-06 refers to Chapter 21 of ACI 318-11 to
shear. Conversely, walls with an aspect ratio of greater than determine the shear strength of existing walls. The nominal
3 are assumed to be slender and their response is consid- shear strength of walls is given in ACI 318-11, Eq. (21-7), as
ered to be controlled by flexure. The response of walls with
intermediate aspect ratios is considered to be controlled by Vn = Acv (a c fc′ + ρt f yt )
both flexure and shear. As the walls presented herein had (1)

286 ACI Structural Journal/March-April 2014


This nominal shear strength is limited to 0.083Acv√fc′ Table 4—Ratios of maximum stress to yield stress
(MPa) (Acv√fc′ [psi]). The coefficient αc is 0.25 for hw/lw ≤ for spliced bars
1.5, is 0.17 for hw/lw ≥ 2.0, and varies linearly between 0.25 φ10 reinforcing bars φ12 reinforcing bars
and 0.17 for intermediate hw/lw values. When determining Test
* †
the yield and nominal shear strengths, ASCE/SEI 41-06 specimen fs/fy fs/fy fs/fy* fs/fy†
limits the strength of the transverse reinforcing bars fyt to the WPS5 0.51 0.80 0.44 0.73
specified yield strength. WPS6 0.51 0.80 0.44 0.73
The nominal shear-friction strength of walls across a
WPS7 0.51 0.80 0.77 1.00
sliding plane perpendicular to shear-friction reinforcing bars
is given by ACI 318-11, Eq. (11-25), as WPS8 0.51 0.80 0.77 1.00
*
Ratio according to Eq. (5).
Vn = (Avf fy + N*)μ (2) †
Ratio according to Eq. (6).

The nominal shear-friction strength given in Eq. (2) is subjected to tension, which is required to be at least 300 mm
limited to the smaller of 0.2Ac fc′ (N) (0.2Ac fc′ [lb]) and 5.5Ac (12 in.), is
(N) (800Ac [lb]).
For the flexural strength of existing walls, ASCE/SEI 41-06  
refers to the basic principles outlined in Chapter 10 of ACI  
fy ψt ⋅ ψe ⋅ ψs ⋅ l 
318-11, but requires the use of expected yield strengths of the ld = 1.3  db (3)
longitudinal reinforcing bars instead of specified minimum  1.1l fc′  cb + K tr  
  d  
yield strengths. When determining flexural yield strengths of  b

walls with no boundary members, ASCE SEI 41-06 requires
considering only the longitudinal reinforcing bars within the
The length of deformed reinforcing bar splices subjected
outer 25% of the wall cross section, but when determining
to compression can be determined from Section 12.16.1 of
the nominal flexural strengths, the contributions of all longi-
ACI 318-11 as
tudinal reinforcing bars within the wall component cross
section need to be considered.
As shown in Table 3, the lateral-force-carrying capacity of  0.071 f y db , f y ≤ 420
ld =   (4)
the test specimens was limited by their flexural strength and,  (0.13 f y − 24)db , f y > 420 
thus, categorizing the response of RC walls by ASCE/SEI
41-06 based on aspect ratio only is found to be misleading.
The results also show that the “plane sections remain plane” This development length is required to be at least 300 mm
hypothesis, which was used during the calculation of the (12 in.) and is required to be increased by one-third for a
flexural strengths, provided strengths that agree well with concrete compressive strength of less than 21 MPa (3.0 ksi).
those determined experimentally, even for squat walls with When the splice length of reinforcing bars within an
an aspect ratio of 1.35. In addition, prior to the peak strength existing wall is found to be inadequate, ASCE/SEI 41-06
of the test specimens being attained, the strain gauge read- stipulates the maximum stress that can be developed within
ings from longitudinal reinforcement of most test specimens the spliced bars to be determined as follows
generally varied linearly across the section.
fs = (lb/ld)fy ≤ fy (5)
Required length of splices
Current provisions for the required length of splices are Ratios of the maximum stresses, which can be developed
based on studies conducted to determine bond-slip relation- by spliced plain round bars according to Eq. (5) to the corre-
ships between isolated reinforcing bars and the surrounding sponding yield strengths of the spliced bars, are presented in
concrete. Transfer of force between starter and main rein- Table 4. Although the provided splice lengths were signifi-
forcing bars over the provided splice length involves a cantly less than those required by Eq. (5) to develop the yield
different force transfer mechanism than that occurring strength of the spliced bars, the walls were able to develop
between an isolated reinforcing bar and the surrounding 97 to 102% of their computed flexural strength (refer to
concrete, but it is widely accepted that the required length Table 3), which was determined assuming that the lap splices
of splices is the same as the required development lengths would develop the yield strength of the spliced bars. Peak
of single embedded reinforcing bars.24 Accordingly, ASCE/ strengths reached by the test specimens with lap-spliced
SEI 41-06 refers to the provisions for required length of reinforcing bars were underestimated by an average of 41%
tension splices of ACI 318-11, which are based on require- (refer to Table 5) by predictions made using Eq. (5). Similar
ment for tension development length. ASCE/SEI 41-06 findings were previously reported in studies24,25 under-
specifies that the required splice length of plain round rein- taken to investigate the behavior of columns with short lap
forcing bars to be taken as twice that required for deformed splices. The lateral force capacity of columns investigated
reinforcing bars. According to ACI 318-11, Eq. (12-1), the by Cho and Pincheira24 was underestimated using Eq. (5)
required length of Class B deformed reinforcing bar splices by an average of 28%. Similarly, columns tested by Melek
and Wallace25 were reported to have achieved 97 to 103%

ACI Structural Journal/March-April 2014 287


Table 5—Measured and calculated strengths of test specimens with spliced longitudinal reinforcement
Test specimen VTest, kN Vn,F-A/S*, kN Vn,F-PS†, kN Vn,F‡, kN VTest/Vn,F-A/S VTest/Vn,F-PS VTest/Vn,F
WPS5 199 91 129 195 2.14 1.51 1.02
WPS6 194 91 144 197 2.16 1.37 0.98
WPS7 231 181 208 233 1.29 1.12 0.99
WPS8 271 234 262 278 1.19 1.06 0.97
Average 1.70 1.27 0.99
Standard deviation 0.46 0.18 0.02

Nominal flexural capacity according to ASCE/SEI 41-06.

Nominal flexural capacity according to proposed supplement26 to ASCE 41-06.
*
Nominal flexural capacity according to ACI 318-11 and assuming splices would develop yield strength of spliced bars.
Note: 1 kN = 0.225 kip.

Fig. 7—Normalized maximum bond stresses that developed between spliced bars and surrounding concrete.
of their yield strengths, which were calculated assuming The maximum stresses that were measured during testing
that the lap splices would develop the yield strengths of the were converted to maximum bond stresses as follows
spliced bars, but the provided splice lengths were approxi-
mately 67% of that required by ASCE/SEI 41-06. f s db
u= (7)
Because Eq. (5) was found to be excessively conserva- 4ld
tive, the provided splice lengths used in the study reported
herein were also compared with those implied by a proposed
supplement26 to ASCE/SEI 41-06. The proposed equation, Using Eq. (3) and (7), the ASCE/SEI 41-06 implied
which is a modified version of Eq. (5) and based on the work maximum bond stress that was expected to develop between
of Cho and Pincheira,24 is a plain round bar subjected to tension and the surrounding
concrete was determined as
0.67
l 
fs = 1.25  b  fy ≤ fy (6) 1  1 1.1 fc′ cb + K tr 
 ld  u=   (8)
2  4 ψ t ψ e ψ s l db 

In most cases, the provided splice lengths were shorter
than required by Eq. (6) (refer to Table 4) to develop the full Maximum bond stresses developed between the spliced
strength of the spliced bars. Predictions made using Eq. (6) bars and the surrounding concrete are compared in Fig. 7,
underestimated the peak strengths by an average of 21%. with maximum bond stress values implied by ASCE/SEI
As discussed previously, both Eq. (5) and Eq. (6) predict 41-06 and the proposed supplement to ASCE/SEI 41-06.
slip to occur at force levels less than the yield strength of the The ASCE/SEI 41-06 implied average maximum bond
spliced bars. However, during testing, all of the monitored stress of 0.29√fc′ (MPa) (3.49√fc′ [psi]) is significantly less
splices developed tensile stresses that were greater than the than the measured average bond stress of 0.57√fc′ (MPa)
experimentally determined yield strength of the spliced bars. (7.95√fc′ [psi]). The average maximum bond strength
implied by the proposed supplement to ASCE/SEI 41-06,

288 ACI Structural Journal/March-April 2014


which was dependent on the provided splice lengths, was seismic behavior of RC walls constructed before the intro-
0.46√fc′ (MPa) (5.54√fc′ [psi]). duction of seismic design requirements in the New Zealand
The performance of the splices reported herein was better Standard Model Building By-law.14 The results of the exper-
than had been expected, principally because of the exces- imental tests were evaluated and compared with current
sively conservative requirement of current design codes19,20 assessment provisions.
and assessment recommendations16 for the required splice Based on the study presented herein, the following conclu-
length of plain round bars. For example, ACI 318 requires sions are drawn:
the length of splices to be increased by one-third if all of 1. The lateral-force-carrying capacity of lightly reinforced
the longitudinal bars are spliced at the same location. This existing walls is limited by their flexural strength. Owing to
requirement is not based on strength criteria,27 but is primarily the low quantity of reinforcing bars provided, yielding of
intended to encourage designers to stagger bar splices. This the longitudinal reinforcing bars dictates the strength of this
requirement alone results in underestimating the capacity of wall type.
spliced reinforcing bars in existing structures by 23%. 2. The strength and the stiffness of this type of wall
In some cases encountered herein, the compression splice degrade rapidly and significantly. The walls have limited
length requirements of ACI 318 governed the required energy dissipation capacity, principally due to the provision
length of splices. The compression splice length require- of few longitudinal reinforcing bars. In addition, the walls
ments of ACI 318 have remained essentially the same since suffer from a lack of transverse confinement reinforcement
the 1963 version of the code and appear to be conservative. to contain concrete in compression zones and to prevent
Required tension splice lengths are expected to govern the longitudinal reinforcing bars from buckling.
required length of splices, as the formation of transverse 3. During testing, the wall pier specimens having no axial
tension cracks around a splice that is subjected to tension load exhibited cracks that were wide during low-level drift
reduces bond strength and, thus, increases the length that is cycles, but the cracks, which were located near the supports,
required to allow the splice to transfer the desired magnitude closed up and appeared inconspicuous after the tests were
of stress. Chun et al.28 have recently reported a better perfor- completed, with longitudinal reinforcement that had previ-
mance of compression splices than tension splices, which ously yielded in tension hidden in the cracks. This type of
was principally attributed to end bearing of reinforcing bars crack could easily be overlooked and the walls may appear
when subjected to compression. Similarly, Cairns29 has intact during post-earthquake inspections, even if the stiff-
found the equations contained in ACI 318-11 to be conserva- ness and the strength of the walls deteriorated significantly.
tive, and has proposed that the length of compression splice 4. From peak stresses measured during testing, it was
for deformed bars be taken as 30% shorter than that required shown that the provisions contained in ASCE/SEI 41-06
by the equations. significantly underestimate the maximum stresses that can
In addition to the aforementioned excessively conservative be developed by plain round reinforcing bar lap splices. The
requirements of ACI 318 for deformed bars, ASCE 41-06 relatively recent recommendations of a proposed supple-
recommends taking the required splice length of plain round ment to ASCE/SEI 41-06 also underestimate the maximum
bars as twice that required for deformed bars, which results stresses that can be developed, but to a lesser extent.
in significantly underestimating the capacity of plain round 5. During testing, the lap splices were able to develop bond
splices located in existing structures. stresses that were significantly higher than the maximum
The good performance of the splices reported herein possible bond stresses implied by ASCE/SEI 41-06. Further
was not considered to be influenced by the thick concrete research is recommended, as the provisions of ASCE/SEI
cover that the splices were provided with. ACI 318 employs 41-06 for required splice lengths of plain round reinforcing
parameters for concrete cover/reinforcement spacing cb bars are based on studies conducted for deformed bars. The
and transverse reinforcement index Ktr to account for provisions are excessively conservative and potentially lead
the confinement term (cb + Ktr)/db, and in ACI 318, it is to unnecessary or expensive seismic retrofitting solutions.
assumed that an increase in the value of this confinement The provisions also lead to the potential for overlooking the
term, above the maximum allowed 2.5, is not likely to danger of existing walls failing in shear during an earthquake,
increase anchorage capacity and is not likely to prevent a before the actual strength of the tension splices is exceeded.
pullout failure, which is a failure mode typically exhibited
by plain round bars. Similarly, Eligehausen et al.30 reported ACKNOWLEDGMENTS
that an increase in concrete cover, reinforcement spacing, or The authors would like to gratefully acknowledge the financial support
provided by the New Zealand Foundation for Research, Science and Tech-
transverse confinement can prevent concrete splitting fail- nology (FRST) through Grant UOAX0411.
ures only, which is a failure mode typically sustained when
using spliced deformed bars. The maximum allowed value AUTHOR BIOS
of (cb + Ktr)/db = 2.5 was employed when calculating the Adane Gebreyohaness is a Structural Engineer at Beca Limited in New
ASCE/SEI 41-06 implied bond strength of the plain round Zealand. His research interests include the seismic assessment, strength-
ening, and design of reinforced concrete and steel structures.
bars discussed herein.
Charles Clifton is an Associate Professor at the University of Auckland,
CONCLUSIONS Auckland, New Zealand. His research interests include the performance
of steel and composite steel/concrete buildings in severe earthquakes and
Twelve test specimens replicated from wall segments of severe fires.
an existing building were experimentally tested to assess the

ACI Structural Journal/March-April 2014 289


John Butterworth is an Associate Professor at the University of Auckland. 7. New Zealand Police, “List of Deceased,” http://www.police.govt.nz/
His research interests include nonlinear structural and solid mechanics; list-deceased. (last accessed Jan. 12, 2012)
stability; buckling behavior of thin-walled sections; structural dynamics; 8. IPENZ, “Christchurch Earthquake—An Overview,” http://www.ipenz.
earthquake engineering; base isolation using rolling, sliding, and rocking org.nz/ipenz/forms/pdfs/ChChFactSheets-Overview.pdf. (last accessed
mechanisms; passive control of structure response (especially by energy Aug. 25, 2011)
dissipating joints); experimental dynamics; pounding of bridges and build- 9. Beca Carter Hollings & Ferner Ltd (Beca), “Investigation into the
ings; and assessment and retrofit of steel structures. Collapse of the Pyne Gould Corporation Building on 22nd February 2011,”
Prepared for Department of Building and Housing (DBH), 2011, 51 pp.
ACI member Jason Ingham is an Associate Professor and Deputy Head 10. Orakcal, K.; Massone, L. M.; and Wallace, J. W., “Shear Strength
(Research) of the Department of Civil and Environmental Engineering at of Lightly Reinforced Wall Piers and Spandrels,” ACI Structural Journal,
the University of Auckland. His research interests include seismic assess- V. 106, No. 4, July-Aug. 2009, pp. 455-465.
ment, retrofit and design of reinforced and prestressed concrete structures, 11. Kuang, J. S., and Ho, Y. B., “Seismic Behavior and Ductility of
and sustainable concrete technology. Squat Reinforced Concrete Shear Walls with Nonseismic Detailing,” ACI
Structural Journal, V. 105, No. 2, Mar.-Apr. 2008, pp. 225-231.
12. Ireland, M.; Pampanin, S.; and Bull, D. K., “Experimental Investi-
gations of a Selective Weakening Approach for the Seismic Retrofit of RC
NOTATION Structural Walls,” NZSEE Conference, Palmerston North, New Zealand,
Ab = area of boundary reinforcement Mar. 30-Apr. 1, 2007, 8 pp.
Ac = area of concrete section resisting shear transfer 13. Greifenhagen, C., and Lestuzzi, P., “Static Cyclic Tests on Lightly
Acv = shear area Reinforced Concrete Shear Walls,” Engineering Structures, V. 27, No. 11,
Al = area of longitudinal reinforcing bar 2005, pp. 1703-1712.
A t = area of transverse reinforcing bar 14. NZ Standards Institute, “NZSS 95: New Zealand Standard Model
Avf = area of shear friction reinforcement Building By-Law, Sections I to X,” NZ Standards Institute, Wellington,
cb = smaller of: (a) distance from center of bar or wire to nearest New Zealand, 1935, 39 pp.
concrete surface; and (b) one-half the center-to-center spacing 15. Gebreyohaness, A. S.; Clifton, G. C.; and Butterworth, J. W.,
of bars or wires being developed “Assessment of Soil-Foundation-Structure Interaction Effects on the
db = nominal bar diameter Seismic Performance of an Old Dual Wall-Frame Building,” 14th ECEE
fc′ = specified compressive strength of concrete Conference, Ohrid, Macedonia, Aug. 30-Sept. 4, 2010, 8 pp.
fs = splice strength 16. ASCE/SEI 41, “Seismic Rehabilitation of Existing Buildings,”
fult = ultimate strength of reinforcement American Society of Civil Engineers, Reston, VA, 2007, 428 pp.
fy = yield strength of reinforcement 17. Oyarzo-Vera, C.; McVerry, G.; and Ingham, J. M., “Seismic Zonation
fyt = yield strength of shear reinforcement and Default Suite of Ground-Motion Records for Time-History Analysis in
hw = height of wall the North Island of New Zealand,” Earthquake Spectra, V. 28, No. 2, 2012,
Ktr = transverse reinforcement index pp. 1-22.
l = length of splice 18. DBH, “Building Act 2004,” Department of Building and Housing,
lb = provided splice length Wellington, New Zealand, 2004, 353 pp.
ld = required splice length according to ACI 318-11 19. ACI Committee 318, “Building Code Requirements for Structural
lw = horizontal length of wall Concrete (ACI 318-11) and Commentary,” American Concrete Institute,
M/Vlw = shear span-to-depth ratio Farmington Hills, MI, 2011, 503 pp.
N* = design axial load 20. NZS 3101:2006, “Concrete Structures Standard: Part 1—The
u = bond stress Design of Concrete Structures,” Standards New Zealand, Wellington, New
Vn = nominal shear strength Zealand, 2006, 696 pp.
VTest = peak strength of test specimen 21. Paulay, T., “Seismic Response of Structural Walls: Recent Devel-
Vu = factored shear force opments,” Canadian Journal of Civil Engineering, V. 28, No. 6, 2001,
λ = modification factor related to density of concrete pp. 922-937.
μ = coefficient of friction 22. ACI Innovation Task Group 1 and Collaborators, “Acceptance
ρl = ratio of area of distributed longitudinal reinforcement to gross Criteria for Moment Frames Based on Structural Testing (ACI T1.1-01),”
concrete area perpendicular to that reinforcement American Concrete Institute, Farmington Hills, MI, 2001, 10 pp.
ρt = ratio of area of distributed transverse reinforcement to gross 23. Brunsdon, D. R., “Seismic Performance Characteristics of Buildings
concrete area perpendicular to that reinforcement Constructed between 1936 and 1975,” University of Canterbury, Christ-
ψe = modification factor based on reinforcement coating church, New Zealand, 1984.
ψs = modification factor based on reinforcement size 24. Cho, J. Y., and Pincheira, J. A., “Inelastic Analysis of Reinforced
ψt = modification factor based on reinforcement location Concrete Columns with Short Lap Splices Subjected to Reversed Cyclic
Loads,” ACI Structural Journal, V. 103, No. 2, Mar.-Apr. 2006, pp.
REFERENCES 280-290.
1. NZSEE, “Assessment and Improvement of the Structural Performance 25. Melek, M., and Wallace, J. W., “Cyclic Behavior of Columns with
of Buildings in Earthquake: Recommendations of a NZSEE Study Group Short Lap Splices,” ACI Structural Journal, V. 101, No. 6, Nov.-Dec. 2004,
on Earthquake Risk Buildings,” New Zealand Society for Earthquake Engi- pp. 802-811.
neering Inc., Wellington, New Zealand, 2012, 343 pp. 26. Elwood, K. J.; Matamoros, A. B.; Wallace, J. W.; Lehman, D. E.;
2. FEMA, “Pre-standard and Commentary for the Seismic Rehabilita- Heintz, J. A.; Mitchell, A. D.; Moore, M. A.; Valley, M. T.; Lowes, L. N.;
tion of Buildings (FEMA 356),” Federal Emergency Management Agency, Comartin, C. D.; and Moehle, J. P., “Update to ASCE/SEI 41 Concrete
Washington, DC, 2000, 519 pp. Provisions,” Earthquake Spectra, V. 23, No. 3, 2007, pp. 493-523.
3. Wood, S. L.; Wight, J. K.; and Moehle, J. P., “The 1985 Chile Earth- 27. ACI Committee 408, “Bond and Development of Straight Reinforce-
quake—Observations on Earthquake-Resistant Construction in Vina Del ment in Tension (ACI 408R-03),” American Concrete Institute, Farmington
Mar,” Structural Research Series No. 532, University of Illinois at Urba- Hills, MI, 2003, 49 pp.
na-Champaign, Champaign, IL, 1987, 192 pp. 28. Chun, S. C.; Lee, S. H.; and Oh, B., “Compression Lap Splice in
4. Sezen, H.; Whittaker, A. S.; Elwood, K. J.; and Mosalam, K. M., Unconfined Concrete of 40 and 60 MPa (5800 and 8700 psi) Compres-
“Performance of Reinforced Concrete Buildings during the August 17, 1999 sive Strengths,” ACI Structural Journal, V. 107, No. 2, Mar.-Apr. 2010,
Kocaeli, Turkey Earthquake, and Seismic Design and Construction Practise pp. 170-178.
in Turkey,” Engineering Structures, V. 25, No. 1, 2003, pp. 103-114. 29. Cairns, J., “Strength of Compression Splices: A Reevaluation of Test
5. EERI, “EERI Special Earthquake Report: The Mw 8.8 Chile Earth- Data,” ACI Journal, V. 82, No. 4, July-Aug. 1985, pp. 510-516.
quake of February 27, 2010,” Earthquake Engineering Research Institute, 30. Eligehausen, R.; Popov, E. P.; and Bertero, V. V., “Local Bond
Oakland, CA, 2010, 20 pp. Stress-Slip Relationships of Deformed Bars under Generalized Exci-
6. Kam, W. Y., and Pampanin, S., “The Seismic Performance of RC tations,” University of California, Berkeley, Berkeley, CA, 1983, 180 pp.
Buildings in the 22 February 2011 Christchurch Earthquake,” Structural
Concrete, V. 12, No. 4, 2011, pp. 223-233.

290 ACI Structural Journal/March-April 2014


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 111-S25

Behavior of Epoxy-Injected Diagonally Cracked Full-Scale


Reinforced Concrete Girders
by Matthew T. Smith, Daniel A. Howell, Mary Ann T. Triska, and Christopher Higgins
Many cast-in place reinforced concrete deck-girder (RCDG) sive bond layer irrespective of the loading conditions (shear,
bridges in the national inventory contain diagonal cracks typically flexure, or tension) (Tremper 1960).
associated with shear-moment interaction. Many transportation As epoxy materials became more widely used, researchers
agencies use epoxy injection as a prophylactic to seal cracks. began to investigate the use of epoxy injection for reinforced
The shear-dominated performance of girders treated with epoxy
concrete members. Early tests to evaluate the performance of
injection is not well understood and the existing data regarding
epoxy-injected reinforced concrete beams were performed
performance of members injected with epoxy have been gathered
from reduced-scale test specimens. To provide new data on the by Chung (1975) with 125 x 200 mm (5 x 8 in.) beam spec-
shear performance of girders with epoxy-injected diagonal cracks, imens on a clear span of 2754 mm (9 ft). Longitudinal and
five full-scale RCDG specimens were constructed to reflect 1950s transverse reinforcement were provided, and each specimen
proportions and details. The specimens were loaded to produce was first loaded to failure and then all major cracks exceeding
diagonal cracks, injected with epoxy under varying degrees of 0.08 mm (0.003 in.) were injected with epoxy resin. Chung
axial tension and service loading, and tested to failure. The test observed that epoxy restored the flexural capacity of the
data indicated that epoxy injection resulted in minimal increased failed specimens. Another study conducted by Popov and
shear capacity. However, the epoxy-injected specimens exhibited Bertero (1975) examined the behavior of full-scale and half-
increased load magnitudes prior to crack re-initiation and reduced scale cantilever reinforced concrete beam-column specimens
stirrup stresses at serviceability levels compared to the cracked
injected with epoxy and exposed to reversed cyclic loading.
condition prior to repair.
The steel detailing and member proportions were typical of
Keywords: bridges; cracking; epoxy; full-scale testing; reinforced concrete; large-sized, short-span cantilevers and beam-column subas-
repair; shear. semblages for buildings. Epoxy injection improved the
original shear strength of the specimens, but the specimens
INTRODUCTION AND BACKGROUND exhibited reduced overall stiffness following repair when
Large numbers of cast-in-place reinforced concrete deck- compared with the uncracked condition. At locations where
girder (RCDG) bridges remain in the national inventory and severe reinforcing bar-concrete bond degradation occurred,
are reaching the end of their originally intended design lives. the epoxy did not perform as well. Additional research by
Many of these bridges are exhibiting varying degrees of Chung (1981) noted that epoxy injection of small 200 x
diagonal-tension cracking in the girders and supporting bent 300 x 2000 mm (7.9 x 11.8 x 78.7 in.) reinforced concrete
caps. Diagonal cracks can be attributed to many sources, beams was not an effective means to restore or improve rein-
including insufficient reinforcing and poor flexural detailing, forcing bar-concrete bond strength.
increasing service load magnitude and volume, and tempera- Basunbul et al. (1990) compared several flexural repair
ture or shrinkage strains. These cracks are generally of methods for reinforced concrete beams, including epoxy
concern to engineers due to the nonductile nature of shear injection. Nine epoxy-injected specimens measuring 150 x
failure and limited reserve shear strength. Further, diag- 150 mm (5.9 x 5.9 in.) in cross section and 1250 mm
onal cracks can expose the embedded reinforcing steel to (49.2 in.) in length, with longitudinal and transverse rein-
chlorides, moisture, and oxygen, which promote corrosion, forcement, were loaded to induce varying degrees of flexural
thereby weakening the structure. Several methods exist that damage. Vertical cracks were injected with epoxy resin and
can seal diagonal cracks and possibly restore or increase allowed to cure before each specimen was loaded to failure.
capacity, such as externally bonded steel and carbon fiber The loads required to reinitiate cracking were observed to
materials, applied ferrocement, cement grouting, and epoxy be approximately 20% higher for the injected specimens
injection. Epoxy, in particular, has been widely used for the compared to the original member response.
rehabilitation of concrete structures. In more recent years, the effects of environment and
The first record of epoxy material testing for highway fatigue have been included in studies of epoxy-injected spec-
applications was performed by the California Division of imens. Abu-Tair et al. (1991) investigated concrete beams
Highways Materials and Research Department (Rooney reinforced with transverse and longitudinal steel measuring
1963). The application was focused on highway mainte- 205 x 140 mm (8 x 5.5 in.) in cross section with a 2300 mm
nance patching of damaged roadways and securing reflec-
ACI Structural Journal, Vol. 111, No. 2, March-April 2014.
tive traffic markers. An earlier study using small-sized plain MS No. S-2012-083, doi:10.14359.51686521, was received March 6, 2012, and
reviewed under Institute publication policies. Copyright © 2014, American Concrete
concrete prisms demonstrated that the bond strength of Institute. All rights reserved, including the making of copies unless permission is
epoxy permitted rupture in the concrete and not in the adhe- obtained from the copyright proprietors. Pertinent discussion including author’s
closure, if any, will be published ten months from this journal’s date if the discussion
is received within four months of the paper’s print publication.

ACI Structural Journal/March-April 2014 291


(90.6 in.) span. Seven specimens were loaded to failure in
flexure and then injected with epoxy resin. Several of the
specimens were loaded monotonically to failure, while
others were fatigue loaded at varying magnitudes. Addi-
tionally, three of the specimens (two monotonically and
one fatigue-loaded) were soaked in 38°C (100°F) water to
investigate the effect of water absorption on durability and
post-repair specimen performance. The results of the study
indicated that epoxy injection restored the original strength
and stiffness of the beams regardless of the loading condi-
tions, while the four months of water immersion had insig-
nificant effects on specimen durability, strength, or stiffness.
The preceding limited research on the structural perfor-
mance of epoxy-injected reinforced concrete specimens has
focused on flexural response of reduced-sized specimens.
No data are available for the shear response of epoxy-in-
jected reinforced concrete girders, and few researchers have
investigated loading conditions on the curing and bonding of
epoxy resin. Furthermore, reduced-sized specimens may not
accurately replicate strain fields and behavior of large rein-
forced concrete members. Other important issues to consider
are in-service loading responses and localized behavioral
effects, incorporation of service-induced diagonal cracks,
and the effects of temperature and shrinkage strains on
epoxy-injected member performance.

RESEARCH SIGNIFICANCE
Epoxy injection is widely used for remediation of cracked
RC bridges and other structures, but the efficacy of the
method on shear performance has not been established. An
experimental program was conducted using realistic full-
scale bridge girders constructed with mid-twentieth century
design methods, details, and materials. Diagonal cracks were
produced under quasi-static loading. The girders were then
epoxy-injected and tested under different loading conditions
to determine the effects of epoxy injection on structural Fig. 1—Specimen reinforcing details and typical instrumen-
performance. Research results improve the understanding tation placement.
of the behavior of epoxy-injected diagonally cracked RC
girders and help engineers make better decisions regarding
rehabilitation alternatives.

EXPERIMENTAL PROGRAM

Test specimens
Five laboratory specimens were constructed and tested
to characterize the behavior and capacity of 1950s vintage
reinforced concrete deck girders with diagonal cracks after
being injected with epoxy resin. Previous work by Higgins et Fig. 2—Specimen naming convention.
al. (2004) identified standard details, materials, and propor-
tions used in 1950s vintage bridge construction. Specimens specimen naming convention used in the study is illustrated
in the current study used an inverted-T (IT) configuration in Fig. 2.
to place the deck in flexural tension. This arrangement is Longitudinal reinforcing steel consisted of ASTM A615/
representative of negative moment in high-shear locations A615M-05a Grade 420 No. 36 (Grade 60, No. 11) bars,
near continuous supports such as piers and bent caps. Each while transverse reinforcement consisted of Grade 300 No.
specimen had the following geometry: 1219 mm (48 in.) 13 (Gr. 40, No. 4). Intermediate grade steel, with a yield
overall height, a stem width of 356 mm (14 in.), and a flange stress of 300 MPa (43.5 ksi), was typically used in 1950s
of 152 mm (6 in.) thick by 914 mm (36 in.) wide. Member construction. However, this grade is not readily available
proportions and reinforcing steel are illustrated in Fig. 1. The for large-diameter reinforcing bars. Therefore, Grade 240
(Grade 60) flexural bars were used but the area was reduced

292 ACI Structural Journal/March-April 2014


Table 1—Concrete and steel material properties
Reinforcing steel
Concrete No. 13, Grade 300 No. 36, Grade 420
fct at initial
Specimen fc′ at failure, MPa (psi) loading, MPa (psi) fy, MPa (ksi) fu, MPa (ksi) fy, MPa (ksi) fu, MPa (ksi)
3.2 (460) at
28.5 (4130) 350 (50.7) 544 (78.9) 477 (69.2) 712 (103)
1-C 28-day
2-EC 36.2 (5250) 2.8 (410) 492 (71.4) 741 (107)
3-ED 28.3 (4104) 2.6 (377) 484 (70.2) 728 (106)
357 (51.8) 570 (82.7)
4-EL 29.5 (4279) 2.6 (377) 473 (68.6) 694 (101)
5-EA 35.4 (5141) 2.8 (410) 492 (71.3) 741 (108)

Note: Values are for failure test unless otherwise indicated.

to produce the same tension resultant as the original designs cally about the midspan. Load was applied in incremental
that used Grade 300 (Grade 40) steel. This results in a steps followed by unloading, and repeating cycles until
smaller area of dowel steel in the specimens compared to the failure. Load magnitudes increased each cycle by 222 kN
original designs. Tension tests were performed according to (50 kip). At each load peak, the load was reduced by 10%
ASTM E8-04 to determine the reinforcing steel properties, to minimize creep effects while visible cracks were marked
which are summarized in Table 1. and recorded.
Concrete was provided by a local ready-mixed concrete Three tests were performed on each specimen: precrack,
supplier. The concrete mixture design was based on 1950s baseline, and failure (with the exception of the control spec-
AASHO “Class A” concrete (Higgins et al. 2003). Speci- imen, 1-C, which was loaded to failure in a single test). An
fied compressive strength was 21 MPa (3000 psi), which is initial loading sequence, or precrack test, was performed
comparable to the specified design strength in the original to produce diagonal cracks similar to those observed from
1950s bridges. Actual concrete compressive strengths were field inspections of RCDG bridges and of sufficient size for
determined from 152 x 305 mm (6 x 12 in.) cylinders tested epoxy injection. A target diagonal crack range of 0.65 to
for 28-day and day-of-test strengths in accordance with 1.25 mm (0.025 to 0.05 in.) was selected based on the earlier
ASTM C39M/C 39M-05 and ASTM C617-05. Day-of-test work of Higgins et al. (2004). When diagonal cracks reached
concrete cylinder strengths for each specimen are shown in a suitable size, the precrack loading cycle was terminated
Table 1. Split cylinder tests in accordance with ASTM C496/ and a baseline test was performed to establish a reference
C496M-04e1 were conducted the same day each specimen for the specimens in the cracked condition for comparison to
was precracked, as reported in Table 1. the post-injection response. The baseline test of the cracked
specimens used the same loading sequence as the precrack
Instrumentation test described previously. In the failure test, specimens were
Internal and external sensors were positioned on the spec- loaded to failure using the same load steps as the previous
imens to record the local and global member responses. two tests.
Strain gauges were placed at midheight on the stirrups
located within the critical shear section near midspan. Addi- Injection and curing procedures
tional strain gauges were mounted to the flexural reinforce- The epoxy resin selected is a commercial, two-part, ultra-
ment at midspan in the flexural-tension region. Diagonal low-viscosity liquid epoxy. The specified material tensile
displacement sensors were placed at three locations along strength is 55.2 MPa (8000 psi). The surface sealant used
the shear span, as shown in Fig. 1. Midspan displacement is a commerical, two-part, 100% solids epoxy. These two
and support settlements were also measured with additional materials are commonly used, are preapproved for use by
displacement sensors. The actual centerline displacement transportation agencies in several states, and are representa-
presented in subsequent figures was calculated by removing tive of similar epoxy materials. All injection materials were
the support deformations from the overall centerline defor- provided by local suppliers. Additional installation guidance
mation during each load cycle. Typical instrumentation is was provided by qualified contractors to establish a repair
illustrated in Fig. 1. protocol that satisfied the installation recommendations of
the manufacturer. The procedure that was established is
Testing methodology summarized as follows.
A simply-supported four-point loading configuration The concrete surfaces around the diagonal cracks were
was used with a span length of 6604 mm (260 in.) from the cleaned with a wire brush and vacuumed to remove loose
centerline of supports. Force was applied from a hydraulic particles and dirt. The crack perimeter was sealed with the
actuator at a quasi-static rate of 8.9 kN/s (2.0 kip/s) and surface epoxy and injection ports were surface-mounted
was measured by a 2224 kN (500 kip) capacity load-cell. A every 356 mm (14 in.), or roughly equal to the width of the
spreader beam distributed the applied actuator force to 102 girder web. The surface epoxy cured for 24 hours before the
mm (4 in.) wide plates spaced 610 mm (24 in.) symmetri- injection process was initiated. To allow for the release of

ACI Structural Journal/March-April 2014 293


Table 2—Specimen experimental summary
Specimen VINITIAL, kN (kip) VAPP, kN (kip) VDL, kN (kip) VEXP, kN (kip) VR2K, kN (kip) VEXP/VR2K Failure mode
1-C N/A 902 (203) 16.9 (3.80) 919 (207) 952 (214) 0.97 Shear-compression
2-EC 723 (162) 983 (221) 20.3 (4.56) 1001 (225) 983 (221) 1.02 Shear-compression
3-ED 778 (175) 992 (223) 17.2 (3.87) 1009 (227) 965 (217) 1.05 Shear-compression
4-EL 778 (175) 1046 (235) 16.7 (3.75) 1063 (239) 947 (213) 1.12 Shear-compression
5-EA 778 (175) 1112 (250) 18.4 (4.14) 1130 (254) 943 (212) 1.20 Shear-compression
Mean 1.07
Coefficient of variation 0.08

entrapped air, diagonal cracks were injected starting from to determine the live load magnitude. Force was applied at
the lowest port working up the crack. “Window” ports were 0.3 Hz with an amplitude of 160 kN (36 kip) and a mean
placed on the backside of the beam to serve as a visual aid of 463 kN (104 kip). This loading amplitude represents the
for assurance of epoxy penetration through the beam web. maximum girder shear caused by the dead load of the bridge
A specialized injection machine, commonly used by local and the truck live load with impact, as well as a minimum
contractors, was needed to mix the two-part liquid epoxy force resulting from hogging due to live load moving onto
in the proportions recommended by the manufacturer and an adjacent span. The loading rate represents the truck trav-
to deliver the mixture into the beam under low pressure. eling over the prototype bridge at approximately 33.8 kph
Each port was injected to a maximum pressure of 690 kPa (21 mph), which was controlled by the hydraulic loading
(100 psi). As liquid epoxy began to seep from the next system in the laboratory.
higher injection port, the lower port was capped and the The fifth specimen, 5-EA, had simulated locked-in drying
injection nipple was moved to the next position. Near the and thermal shrinkage strains induced by applying a uniform
top of each diagonal crack, the injection pressure climbed tension load to the specimen. The axial load was applied
more quickly to 690 kPa (100 psi), and would take longer to to a level of approximately 890 kN (200 kip) before the
dissipate, signaling that there was little available void space initial precrack transverse loading cycles began. The axial
to pump additional resin. When the maximum pressure was load was held at a constant magnitude of 645 kN (145 kip)
maintained, the final port was capped. After injection, all during the injection and curing phases, and then returned to
specimens were allowed to cure for at least 7 days. A heated 890 kN (200 kip) for the post-injection failure loading. For
plastic enclosure was placed around the specimen to ensure additional detail on the axial force application and loading
that temperatures were maintained above 4.5°C (40°F), as protocol, refer to Smith (2007).
recommended by the epoxy manufacturers. Thermocouples
outfitted with data loggers were placed into a small void cast EXPERIMENTAL RESULTS
into the end of the specimens and on the exterior to record The performance of the epoxy-injected specimens was
temperatures throughout the curing cycle. evaluated through the shear-midspan deflection and shear-
diagonal displacement responses, flexural and shear rein-
Specimen variables forcement strains, and crack deformations. The data
To simulate the effects of different in-service stress condi- collected from the three phases of testing were compared to
tions, each specimen was subjected to a distinct loading assess the local and global responses before and after epoxy
scenario during the injection and curing phases. Specimen injection. Results were also compared with an otherwise
2-EC was injected and cured with no applied loads other similar un-injected specimen. All of the specimens exhibited
than specimen self-weight. Simulated superstructure dead shear-compression failures and the applied shear at failure
load was applied to Specimen 3-ED before epoxy injec- for each specimen is summarized in Table 2.
tion. A total load of 356 kN (80 kip) was applied to induce a
service level dead load shear of 178 kN (40 kip). This shear Shear-midspan displacement response
magnitude is representative of an interior girder for a typical Applied shear-midspan displacement responses are shown
1950s vintage three-span continuous RCDG bridge having in Fig. 3 and demonstrate the overall specimen behavior of
15.2 m (50 ft) spans and a uniform dead load of 23.3 kN/m/ the initial, baseline, and post-injection tests. The post-injec-
girder (1.6 kip/ft/girder). tion response of each injected specimen showed decreased
Varying live load stress was applied to Specimen 4-EL in residual deformations and greater stiffness during the first
addition to the superstructure dead load. The live loading two or three load steps as compared to the baseline response.
was representative of average shear magnitudes produced by As the applied shear magnitudes increased, the specimens
ambient traffic and a fully loaded AASHTO Type 3-3 unit began to soften due to the development of new cracks, often
truck having five axles and a gross vehicular weight of 356 adjacent to the repaired diagonal cracks. Specimens 3-ED
kN (80 kip) moving across the bridge described for Spec- and 4-EL were similar, especially in the service load range
imen 3-ED. Realistic shear distribution factors developed indicated in Fig. 3, and both had greater stiffness than Spec-
by Potisuk and Higgins (2007) from field studies were used imen 2-EC.

294 ACI Structural Journal/March-April 2014


Fig. 3—Applied shear-midspan displacement response.
shear decreased, the specimen shortened along the axis of
the axial loading apparatus and the axial tension increased
again. Like Specimen 4-EL, Specimen 5-EA had decreased
residual deformations for many of the load cycles and did
not begin to soften until near failure.

Shear-diagonal displacement response


The post-injection diagonal displacement data are shown
in Fig. 5. All of the post-injection tests for the epoxy speci-
mens had greater stiffness and smaller permanent deforma-
tions during the initial load steps than the control specimen.
Specimens 3-ED and 4-EL were stiffer than Specimen 2-EC,
with Specimen 4-EL performing slightly better than 3-ED.
Specimen 5-EA showed significantly improved stiffness and
Fig. 4—Applied shear-axial load variability of Specimen reduced permanent deformations after unloading following
5-EA. epoxy injection. For all the specimens except 4-EL and
5-EA, the north and south diagonal deformations were
The axially loaded specimen exhibited a unique shear-mid- essentially identical, showing similar stiffness and exhib-
span displacement response. The curve has a slender S-shape iting increasing diagonal deformation at approximately the
resulting from the specimen stiffening and then softening same shear magnitude. For the remaining two specimens,
during each load cycle. The axial load as a function of the however, one end of the specimen produced a substantially
applied shear for Specimen 5-EA is displayed in Fig. 4. As larger diagonal deformation than the other side.
the applied shear increased, the specimen length increased
at the level of the axial apparatus, thereby reducing the Diagonal cracking behavior
hydraulic pressure in the axial load actuators, and thus The orientation and location of the diagonal cracks
reducing the externally applied axial tension. As the applied produced during precrack and post-injection loading

ACI Structural Journal/March-April 2014 295


Fig. 5—Control and post-injection shear-diagonal displacement response recorded near centerline of specimens.
sequences are shown in Fig. 6. The locations of diagonal Reinforcement strains
cracks that were epoxy-injected are also shown in Fig. 6. The largest relative influences of epoxy injection were
Diagonal cracks that were injected did not reopen during seen in the individual stirrup strains, but these effects were
post-injection tests. Instead, new cracks formed adjacent to highly influenced by the proximity of injected diagonal
the injected cracks and propagated at similar angles. Non-re- cracks to the embedded strain gauge locations. Strain gauges
paired cracks propagated along the original paths. located near diagonal cracks that were injected had lower
The applied loads required to reinitiate diagonal cracking strains after injection at similar shear magnitude. Strain
are shown in Fig. 7. The load required to reinitiate diagonal gauges located between diagonal cracks or far from cracks
cracking was determined from the applied shear magnitude that were not injected displayed relatively little change. This
at the moment the stirrup strain showed an abrupt increase. behavior was observed for all injected specimens. Diagonal
The pre- and post-injection diagonal cracking shears were cracks were considered to be “near” the stirrup strain gauge
compared with the precrack diagonal cracking shear on the if the vertical distance that the crack crossed the stirrup with
abscissa and the post-injection cracking shear serving on the strain gauge was within the AASHTO-LRFD (AASHTO
the ordinate. Solid symbols represent stirrup strain gauges 2005) calculated development length of the Grade 300 No. 13
located near injected diagonal cracks, while hollow symbols (Grade 40, No. 4) stirrup (203 mm [8 in.]). An example of
represent stirrup strain gauges located away from injected strains measured for a stirrup located near a diagonal crack
diagonal cracks, as described in the next section. Injected and a stirrup located at a distance greater than the develop-
diagonal cracks required higher applied shear than the orig- ment length from a diagonal crack is shown in Fig. 8.
inal specimen to produce strains in the stirrups due to new The strain behavior depicted in Fig. 9 shows the baseline
diagonal cracking. Data above the reference line show that and post-injection stirrup strains at the maximum service
larger shear loads were required, while points below the line load range. In the figure, the baseline strains serving as the
required smaller loads to propagate or reinitiate diagonal abscissa are plotted against the post-injection strains on
cracking. Non-injected cracks typically behaved similarly the ordinate. Solid symbols represent stirrup strain gauges
to the baseline tests, where stirrup strains began increasing located near injected diagonal cracks, while hollow symbols
immediately upon application of applied shear. represent stirrup strain gauges located away from injected
diagonal cracks, as defined previously. The dashed refer-

296 ACI Structural Journal/March-April 2014


Fig. 6—Crack pattern locations on east face of specimens. Figure includes precracking, epoxy-injected cracks, post-injection
cracks, and final failure crack.
ence line marks the boundary between improved and unim- Interaction of epoxy curing process and cyclic live
proved behavior. Points above the line had higher strains load
at the same service load after injection, whereas the points Data were collected continuously for Specimen 4-EL
below the line had lower strains after injection. Most stirrup during curing of the epoxy with cyclic service-level live
strains near injected diagonal cracks showed significantly load being applied. An example of the diagonal deformation
reduced strains after injection, whereas uninjected regions throughout the first 3 days of curing is shown in Fig. 10. The
were generally unaffected. deformation range of a representative displacement sensor
crossing the injected diagonal crack is shown in Fig. 11.
As seen in Fig. 10, within the first several hours of curing,

ACI Structural Journal/March-April 2014 297


the average diagonal crack deformation reduced from 1.39 stirrup is located in the same section as the example diagonal
to 1.27 mm (0.0547 to 0.0500 in.) and remained relatively deformation shown in Fig. 11, and the strain range decreased
constant for the remainder of the curing process. However, by over 50% within the first 18 hours, while little addi-
the diagonal deformation range continued to reduce over tional effects were observed for the remainder of the curing
the period of about 3 days, during which the deformation process. The curing time reported by the epoxy manufac-
range decreased by nearly 75% (Fig. 11). An example of the turer is 7 days at 4°C (40°F) and 2 days at 25°C (77°F). The
stirrup strain throughout the curing period and the internal average curing temperature for Specimen 4-EL was 10°C
and external temperature recordings is shown in Fig. 12. The (50°F), which correlates to a required curing time of approx-
imately 6 days. The stirrup strain was also observed to fluc-
tuate with the external temperature with significant time lag
between surface temperature and strain changes (Fig. 12).

DISCUSSION
The results of this study indicate that epoxy injec-
tion affected the structural behavior of the RC specimens
in several ways. Overall, the most dramatic effects were
observed for Specimens 3-ED, 4-EL, and 5-EA, as described
in the following discussion.

Shear capacity of specimens


In this study, a computer program call Response 2000
(R2K) (Bentz 2000) was used to estimate the strength of the
specimens (neglecting any influence of the epoxy injection)
as well as the strength of the control specimen. In a previous
study, R2K, which uses the Modified Compression Field
Theory (MCFT) (Vecchio and Collins 1986), was used to
Fig. 7—Applied shear at diagonal cracking initiation before predict shear capacity for a series of 31 similar full-size RC
and after injection. specimens within 0.98 of the actual capacity with a coeffi-

Fig. 8—Applied shear-stirrup strain behavior for Specimen 3-ED. Strain gauges located near: (a) injected diagonal cracks;
and (b) uninjected diagonal cracks.

298 ACI Structural Journal/March-April 2014


Fig. 11—Diagonal deformation range during curing of
Specimen 4-EL. (Note: Zero was taken as a reference value.)
Fig. 9—Stirrup strains at service level shear before and after
injection. Shear magnitude taken as 311 kN (70 kip).

Fig. 12—Stirrup strain during curing of Specimen 4-EL.


(Note: °C = [°F × 1.8] + 32.)
Fig. 10—Diagonal deformations reduced during curing of
Specimen 4-EL. Interaction of epoxy repair and superimposed
dead load
cient of variation under 8% (Higgins et al. 2004). The R2K Specimen 3-ED exhibited a higher capacity than the
predicted member capacities are shown in Table 2. The predicted baseline capacity. Specimen 3-ED also achieved
experimental shear strength is the applied actuator force at higher load prior to reinitiation of nonlinear response
failure combined with the self-weight of the specimen acting compared to the epoxy-injected control specimen, 2-EC. The
at the failed section. Except for two specimens, the epoxy- dead load serves to keep the diagonal cracks open, allowing
injected specimens exhibited slightly larger shear capacities for more penetration of the epoxy at the crack tips. It further
than predicted, ranging from 1.02 to 1.12. However, all but allows the epoxy to only carry superimposed live loads and
4-EL and 5-EA fall within a standard deviation (68% predic- leaves dead load stresses locked into the reinforcement and
tion interval) of the expected shear strength ignoring the concrete. This tends to further delay crack reinitiation. This
effects of epoxy injection. This is within the expected vari- was observed by the reduced stirrup steel demand at injected
ability of shear testing results. Specimens 4-EL and 5-EA diagonal cracks for otherwise similar load levels.
exhibited shear strengths significantly above the predicted
unaltered shear capacity. The shear strength of Specimen Cured epoxy characteristics following cyclic live
4-EL was above the 95% prediction interval, and Specimen load
5-EA was above the 99% prediction interval for the expected The live load magnitudes, rates, and curing conditions
shear strength without epoxy injection and indicates the considered in this program for Specimen 4-EL did not
increase is not likely attributed to the inherent variability of reduce the effectiveness of the epoxy injection compared to
shear strength testing of large reinforced concrete girders of the specimen with dead load alone and, in fact, resulted in
the type studied herein. These two specimens received the higher observed shear capacity. The cyclic loading acted as
least amount of epoxy injection in comparison to the other an internal pumping mechanism that enabled the epoxy to
specimens, because they only exhibited three major diagonal enter and fill finer cracks than in either Specimen 3-ED or
crack systems injected on each specimen. Axial load was Specimen 2-EC. It was observed during the injection process
accounted for in the R2K capacity prediction of Specimen of Specimen 4-EL that the epoxy pump pressures built and
5-EA but not 4-EL, and the differences in shear capacity for dissipated in-phase with the actuator loading cycle, which is
each of the specimens is described subsequently.

ACI Structural Journal/March-April 2014 299


tration of the epoxy into microcracks within the concrete
matrix, similar to Specimens 3-ED and 4-EL. More impor-
tantly, as the vertical loading was applied, the axial force
decreased during the failure test to a magnitude of 267 kN
(60 kip) at ultimate load, resulting in a net compressive force
of 378 kN (85 kip) induced into the epoxy-injected section.
Had the axial load not diminished with increasing transverse
load, the specimen may have failed at lower load levels.
R2K predicted a capacity of 853 kN (192 kip) for a similar
Fig. 13—Photographs of cores taken from Specimens 3-ED specimen with 890 kN (200 kip) total axial tension force.
and 4-EL: (a) core taken from Specimen 3-ED; and (b) core Reducing the steel yield stress by an amount equivalent to
taken from Specimen 4-EL with small voids. the 267 kN (60 kip) axial tension and applying a 378 kN (85
kip) axial compression force on the section, R2K estimated
consistent with the epoxy being pushed out of the diagonal a shear capacity of 987 kN (222 kip), which is closer to the
cracks as the cracks closed upon unloading. observed shear capacity. This situation of loading and curing
Concrete cores measuring 102 mm (4 in.) in diameter would represent a structure with shrinkage and/or tempera-
were taken from epoxy-injected diagonal cracks for both ture-induced tensile strains that are recovered after epoxy
Specimens 3-ED and 4-EL after failure. Both cores showed injection. The strain recovery (release of restraints at supports
that the epoxy was well distributed through the cracks and or temperature change, for example) produces a post-ten-
even filled hairline subcracks within the cored region. The sioning effect for the injected cross section. However, the
core taken from Specimen 4-EL had small visible voids, beneficial temperature effect could not be relied upon in the
which were evidence of bubble formations likely caused by field and would vary during daily and seasonal changes.
the cyclic loading noted previously. Examples of the porous
epoxy matrix observed in Specimen 4-EL compared to the CONCLUSIONS
solid epoxy matrix observed in Specimen 3-ED are shown Five RC deck girder specimens were fabricated to reflect
in Fig. 13. The development of bubbles within the epoxy did the design and construction materials of the 1950s for
not diminish the performance of Specimen 4-EL compared RCDG bridges lightly reinforced for shear. The specimen
to Specimen 3-ED. It is important to note that the cyclic live tests were designed to study the effects of epoxy on diag-
loading was representative of loads moving across a typical onal-tension, shear-dominated cracked girders. Specimens
15.2 m (50 ft) span continuous bridge at an approximate were precracked to similar levels observed in the field,
speed of 33.8 kph (21 mph). Additionally, due to setting of subjected to baseline tests in the cracked condition, injected
the epoxy at a larger average crack width (dead load plus with epoxy resin at varying levels of applied dead and/or
average live load range), as compared to Specimens 2-EC live load, and then, after curing the epoxy, were tested to
and 3-ED, the cross section was effectively induced with failure. The results of the initial cracking, baseline, and post-
compressive stresses that must be overcome before addi- injection responses were compared. Factors included in the
tional cracking may occur. This post-tensioning effect also study were superimposed dead load, service live load plus
accounts for the increased shear strength observed for Spec- dead load, and externally applied axial tension during epoxy
imen 4-EL. Further research is needed to study the effects curing. Based on the experimental observations and analyt-
of higher loading rates and other load magnitudes applied ically predicted shear strengths, the following conclusions
simultaneously during epoxy injection, as well was possible are presented:
lower-range curing temperatures. However, based on these • Most of the epoxy-injected specimens exhibited
observations, it may be possible to inject cracks on existing marginal increases in shear strength compared to
bridges while positioning static superimposed live loads on well-correlated analytically predicted strengths of
the bridge (such as loaded maintenance trucks) to effectively unrepaired specimens. The largest capacity increases
open the cracks to their maximum service-level width. After were observed for specimens subjected to superim-
curing the epoxy and removing the live load, compressive posed cyclic live load and externally applied static axial
stresses would be induced in the cross section. If the live tension during epoxy curing (which was due principally
load used during repair is above the maximum expected to an unintended post-tensioning effect). These showed
service loads, this could effectively prevent or significantly that the epoxy injection did not effectively strengthen
delay future cracking. the specimens, as they would have failed very close to
the observed capacity with or without epoxy injected
Interaction of epoxy repair and axial load cracks.
The axially loaded specimen, 5-EA, exhibited the most • Superimposed cyclic live loading during injection and
dramatic change between the pre- and post-injection curing of Specimen 4-EL produced dynamic pressure
response. The specimen was injected with a constant exter- during injection and pumping of the epoxy within the
nally applied axial tension force of 645 kN (145 kip), which diagonal cracks. Fine bubbles were formed within
coincided with the load magnitude at the end of the precrack the epoxy matrix but did not reduce structural perfor-
test. The applied axial force maintained larger diagonal and mance at service or ultimate states. In addition, the
vertical cracks during injection, allowing for increased pene- curing of the epoxy at larger average diagonal crack

300 ACI Structural Journal/March-April 2014


widths introduced compressive stresses in the cross NOTATION
section that increased the threshold load for recracking. fc′ = compressive strength of concrete, MPa (psi)
fct = tensile strength of concrete, MPa (psi)
The compressive stresses in the stem also resulted in fu = ultimate tensile stress of reinforcing steel, MPa (ksi)
increased shear capacity. Consequently, this finding fy = yield stress of reinforcing steel, MPa (ksi)
may permit strategic placement of loaded maintenance VAPP = applied shear from actuator, kN (kip)
VDL = applied shear from portion of self-weight acting at failure plane,
trucks to widen cracks at certain locations during the kN (kip)
injection and curing process. VEXP = total applied shear, kN (kip)
• Initial stiffness after injection was improved and devel- VINITIAL = maximum applied shear during precracking phase of testing, kN
(kip)
opment of residual deformations was delayed prior VR2K = predicted shear capacity, kN (kip)
to recracking by epoxy injection compared with the
cracked performance prior to injection. REFERENCES
• Epoxy injection increased the threshold load level AASHTO, 2005, “AASHTO-LRFD Bridge Design Specification,” third
required to form additional diagonal cracks or extend edition with 2005 interims, American Association of State Highway and
Transportation Officials, Washington, DC, 654 pp.
the existing cracks within the stem. Abu-Tair, A. I.; Rigden, S. R.; and Burley, E., 1991, “The Effectiveness
• Injected diagonal cracks did not reopen. Instead, new of the Resin Injection Repair Method for Cracked RC Beams,” The Struc-
cracks formed adjacent to the original injected cracks. tural Engineer, V. 69, No. 19, Oct., pp. 335-341.
ASTM A615/A615M-05a, 2005, “Standard Specification for Deformed
• Epoxy injection reduced service-level stirrup strains and Plain Carbon-Steel Bars for Concrete Reinforcement,” ASTM Interna-
compared to uninjected diagonal cracks prior to tional, West Conshohocken, PA, 6 pp.
recracking. This may reduce bond fatigue, thereby ASTM C39/C39M-05, 2005, “Standard Test Method for Compressive
Strength of Cylindrical Concrete Specimens,” ASTM International, West
slowing or preventing additional diagonal crack growth Conshohocken, PA, 7 pp.
and help maintain force transfer across diagonal cracks. ASTM C496/C496M-04e1, 2004, “Standard Test Method for Splitting
• Careful and methodical epoxy installation proce- Tensile Strength of Cylindrical Concrete Specimens,” ASTM International,
West Conshohocken, PA, 5 pp.
dures following industry best practices enabled epoxy ASTM C617-05, 2005, “Standard Practice for Capping Cylindrical
penetration through the depth of the web. Based on Concrete Specimens,” ASTM International, West Conshohocken, PA, 6 pp.
the observed penetration and uniformity of the epoxy ASTM E8-04, 2004, “Standard Test Methods for Tension Testing of
Metallic Materials,” ASTM International, West Conshohocken, PA, 28 pp.
within the diagonal cracks, it is likely that the epoxy Basunbul, I. A.; Gubati, A. A.; Al-Sulaimani, G. J.; and Baluch, M. H.,
would restrict access of moisture and chlorides to the 1990, “Repaired Reinforced Concrete Beams,” ACI Materials Journal,
embedded stirrups and thereby delay or diminish corro- V. 87, No. 4, July-Aug., pp. 348-354.
Bentz, E. C., 2000, “Section Analysis of Reinforced Concrete Members,”
sion potential at the injected crack locations. PhD thesis, Department of Civil Engineering, University of Toronto,
Toronto, ON, Canada.
ACKNOWLEDGMENTS Chung, H. W., 1975, “Epoxy-Repaired Reinforced Concrete Beams,”
This research was funded by the Oregon Department of Transportation. ACI Journal, V. 72, No. 5, May, pp. 233-234.
The findings and conclusions are those of the authors and do not necessarily Chung, H. W., 1981, “Epoxy Repair of Bond in Reinforced Concrete
reflect those of the project sponsors. Members,” ACI Journal, V. 78, No. 1, Jan.-Feb., pp. 79-82.
Higgins, C.; Farrow III, W. C.; Potisuk, T.; Miller, T. H.; Yim, S. C.;
Holcomb, G. R.; Cramer, S. D.; Covino, B. S.; Bullard, S. J.; Ziomek-
AUTHOR BIOS Moroz, M.; and Matthes, S. A., 2003, “SPR 326 Shear Capacity Assess-
Matthew T. Smith is an Associate Engineer at CH2M-Hill, Corvallis, ment of Corrosion-Damaged Reinforced Concrete Beams,” Oregon Depart-
OR. He received his BS and MS from Oregon State University, Corvallis, ment of Transportation, Salem, OR, 19 pp.
OR. His research interests include the design of hydraulic and building Higgins, C.; Miller, T. H.; Rosowsky, D. V.; Yim, S. C.; Potisuk, T.;
structures. Daniels, T. K.; Nicholas, B. S.; Robelo, M. J.; Lee, A.-Y.; and Forrest,
R. W., 2004, “Research Project SPR 350 SR 500-91: Assessment Method-
Daniel A. Howell is a Project Manager with the Bridge Design Division of ology for Diagonally Cracked Reinforced Concrete Deck Girders,” Oregon
the St. Louis County Department of Highways and Traffice, Creve Coeur, Department of Transportation, Salem, OR, Oct., 328 pp.
MO. He received his BS and MS from the University of Delaware, Newark, Popov, E. P., and Bertero, V. V., Oct.-Dec. 1975, “Repaired R/C
DE, and his PhD from Oregon State University. His research interests Members Under Cyclic Loading,” Earthquake Engineering & Structural
include the design of bridges and other structures. Dynamics, V. 4, No. 2, pp. 129-144.
Potisuk, T., and Higgins, C., 2007, “Field Testing and Analysis of CRC
Mary Ann T. Triska is a Bridge Engineer at HDR Engineering, Portland, Girder Bridges,” Journal of Bridge Engineering, V. 12, No. 1, Jan.-Feb.,
OR. She received her BS from the University of Portland, Portland, OR, pp. 53-63.
and her MS from Oregon State University. Her research interests include Rooney, H. A., 1963, “Epoxy Resins as a Structural Repair Mate-
evaluation, rehabilitation, and sustainable design of infrastructure. rial,” State of California Department of Public Works Division of
Highways, Jan., 15 pp. (http://www.dot.ca.gov/hq/research/researchre-
Christopher Higgins is the Slayden Construction Faculty Fellow and ports/1961-1963/63-23.pdf)
Professor of Structural Engineering in the School of Civil and Construc- Smith, M. T., 2007, “Investigation of the Behavior of Diagonally Cracked
tion Engineering at Oregon State University. He received his BS from Full-Scale CRC Deck-Girders Injected with Epoxy Resin and Subjected to
Marquette University, Milwaukee, WI; MS from the University of Texas Axial Tension,” MS thesis, Oregon State University, Corvallis, OR. (http://
at Austin, Austin, TX; and PhD from Lehigh University, Bethlehem, PA. scholarsarchive.library.oregonstate.edu.)
His research interests include evaluation and rehabilitation of bridges and Tremper, B., 1960, “Repair of Damaged Concrete with Epoxy Resins,”
other structures. ACI Journal, V. 57, No. 2, Feb., pp. 173-182.
Vecchio, F. J., and Collins, M. P., 1986, “The Modified Compression
Field Theory for Reinforced Concrete Elements Subjected to Shear,” ACI
Journal, V. 83, No. 2, Feb., pp. 219-231.

ACI Structural Journal/March-April 2014 301


NOTES:

302 ACI Structural Journal/March-April 2014


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 111-S26

High-Performance Fiber-Reinforced Concrete Bridge


Columns under Bidirectional Cyclic Loading
by Ady Aviram, Bozidar Stojadinovic, and Gustavo J. Parra-Montesinos
An experimental and analytical study was carried out on circular substantial experimental evidence that supports the use of
column specimens representing cantilever bridge piers constructed HPFRC materials to enhance structural response of elements
with tensile strain-hardening, high-performance fiber-reinforced subjected to large deformation demands caused by ground
concrete (HPFRC). Two column specimens with different longitu- motions, such as bridge piers with either flexural-dominated
dinal reinforcement details in the expected plastic hinge zone were
behavior or with strong flexure-shear interaction.
tested under bidirectional displacement reversals, and the results
Evaluation of the use of HPFRC in bridge piers to
were compared with those of a geometrically identical specimen
constructed using regular concrete and designed according to substantially relax transverse reinforcement requirements
current Caltrans bridge design specifications. The results demonstrate while leading to increased damage tolerance, shear strength,
the considerable benefits of using tensile strain-hardening fiber- and energy dissipation under cyclic loading compared with
reinforced concrete in typical highway overpass bridge columns. regular concrete piers was the main focus of this research.
These benefits include: improved cyclic response; substantial An experimental and analytical study was carried out on
reduction in transverse reinforcement and corresponding construc- two approximately 1/4-scale column specimens built with
tion benefits; and increased damage tolerance and reduction of HPFRC and subjected to bidirectional displacement rever-
post-earthquake repair costs. sals. The behavior of these specimens was compared with
that of a geometrically identical conventionally reinforced
Keywords: bidirectional cyclic load; bond stress; bridge column; drift;
fiber-reinforced concrete; plastic hinge; shear; steel fibers. concrete (RC) column.6 Additional information about the
tests, calibrated finite element models, and the repair cost and
INTRODUCTION repair time analysis of typical highway bridges in California
Numerous cast-in-place and precast reinforced concrete constructed using HPFRC columns can be found elsewhere.7
structures suffered significant damage or collapse during
historical and recent earthquakes, primarily due to deficient RESEARCH SIGNIFICANCE
structural design. Code-mandated reinforcement detailing Experiments were performed to characterize the cyclic
required for critical bridge and building members to ensure response of circular HPFRC bridge columns subjected to
adequate seismic behavior often leads to substantial rein- bidirectional lateral displacements and to compare their
forcement congestion and construction difficulties. There- response with that of conventionally reinforced concrete
fore, it is not surprising that many recent research efforts columns. This experimental study is one of the first of its
have been directed to the development and implementa- kind, and was aimed at assessing the effectiveness of fiber
tion of innovative materials in new structures for improved reinforcement as partial replacement of transverse rein-
seismic performance while simplifying the required forcement used for shear resistance and confinement while
reinforcement detailing. increasing column flexural ductility. The experiments
Fiber-reinforced concretes that exhibit a tensile strain-hard- performed also allowed an evaluation of the enhanced
ening behavior are now possible with the use of relatively low damage tolerance of HPFRC columns, which is important
fiber-volume fractions Vf (in the range of 1.5 to 2.0%). These for reducing post-earthquake repair cost and repair time
tensile strain-hardening materials are typically referred to as assessment, and performance-based evaluation of structural
high-performance fiber-reinforced concrete (HPFRC). In systems using HPFRC.
addition to their tensile strain capacity, which often exceeds
0.5%, HPFRCs exhibit a compression response that resem- TEST PROGRAM
bles that of well-confined concrete. Hooked and twisted The column specimens represented the bottom half of a
steel fibers, as well as ultra-high-molecular-weight polyeth- typical bridge column deforming in double curvature with
ylene fibers, are among the fiber types investigated for use an assumed inflection point at midheight. The specimen
in earthquake-resistant construction.1 When used in struc- geometry was selected so as to represent circular column
tural elements subjected to large displacement reversals, prototypes used by Caltrans for typical highway overpass
HPFRCs enable significant deformation capacity with supe- bridges in California. The length scale of the specimens was
rior damage tolerance compared with geometrically iden- approximately 1/4. Columns of both HPFRC specimens,
tical, well-detailed reinforced concrete members.1-5 Further, ACI Structural Journal, V. 111, No. 2, March-April 2014.
substantial reductions in transverse reinforcement required MS No. S-2012-092, doi:10.14359.51686522, was received March 10, 2012, and
reviewed under Institute publication policies. Copyright © 2014, American Concrete
for confinement and shear resistance have been possible in Institute. All rights reserved, including the making of copies unless permission is
obtained from the copyright proprietors. Pertinent discussion including author’s
elements subjected to large shear stress reversals.1 There is closure, if any, will be published ten months from this journal’s date if the discussion
is received within four months of the paper’s print publication.

ACI Structural Journal/March-April 2014 303


Table 1—Summary of specimen geometry, reinforcement, and material properties
Parameter S1: HPFRC S2: HPFRC BC: Plain
Column diameter Dcol 400 mm (16 in.) 400 (16 in.) 400 mm (16 in.)
Total column height Hcol 1625 mm (64 in.) 1625 mm (64 in.) 1625 mm (64 in.)
Shear span-to-diameter ratio 4 4 4
12 No. 4/13M + 8 No. 4/13M
Longitudinal reinforcement 12 No. 4/13M + 8 No. 4/13M dowels 12 No. 4/13M
dowels
Longitudinal reinforcement ratio ρl 2% (base), 1.2% (rest) 2% (base), 1.2% (rest) 1.2%
Dowels, L = 250 mm Main reinforcing bar, L = 100 mm (4
Debonding sleeves —
(10 in.), 250 mm (10 in.) above base in.), 200 mm (8 in.) above base
Transverse reinforcement W3.5* at 64 mm (2.5 in.) W3.5 at 64 mm (2.5 in.) W3.5 at 32 mm (1.25 in.)
Volumetric transverse reinforcement ratio ρs 0.75% 0.75% 1.5%
Specified concrete compressive strength 34.5 MPa (5.0 ksi) 34.5 MPa (5.0 ksi) 34.5 MPa (5.0 ksi)
Days from concrete casting to test date 60 49 49
Concrete compressive strength at test day 47.3 MPa (6.86 ksi) 47.1 MPa (6.83 ksi) 35.1 MPa (5.09 ksi)
*
W3.5: diameter = 5.3 mm (0.21 in.).

denoted as S1 and S2 in Table 1, were built using the same Two different debonding schemes, shown in Fig. 1, were
HPFRC material, reinforced with a 1.5% volume fraction of devised to minimize damage localization at the section where
commercially available high-strength (2410 MPa [350 ksi]) the dowel reinforcement was terminated. In Specimen S1,
hooked steel fibers with a length of 30 mm (1.2 in.), and the upper portion of the dowels was debonded using plastic
an aspect ratio of 80. A fiber-volume fraction of 1.5% tubes to avoid premature damage localization that could
(120 kg/m3 [200 lb/yd3]) was selected such that a tensile occur because of the termination of the bars within the
strain-hardening behavior could be achieved. Ready mix plastic hinge zone. In Specimen S2, the dowels were termi-
concrete was used in both columns, and the fibers were added nated within the plastic hinge region, and the main longitu-
to the concrete truck on site. HPFRC casting was performed dinal bars were debonded over a length of 4 in. (10 cm) to
using a bucket and stick vibrator. The results from the test of prevent large strain concentration and premature reinforcing
a reference base column specimen (BC), designed according bar fracture at the section where the dowels were terminated.
to the Caltrans Seismic Design Criteria (SDC)8 and built
with regular concrete, were used for comparison purposes.6 Loading pattern
All three specimens were tested following the same prede- The HPFRC specimens were tested using a bidirectional
termined bidirectional cyclic displacement pattern. circular load pattern (Fig. 2). This pattern was similar to that
The transverse reinforcement of both HPFRC speci- used for the BC specimen.6 A cycle in this pattern comprised
mens was approximately half of that required by Caltrans circles, one clock-wise and the other counterclockwise,
Seismic Design Criteria (SDC),8 counting on the HPFRC while return paths were sequenced to minimize the bias in
material to contribute significantly to shear resistance and any particular loading direction; however, this resulted in
confinement. The longitudinal reinforcement of the columns a very demanding displacement pattern characterized by
in Specimens S1 and S2, shown in Fig. 1, was detailed to long specimen travel along circles with constant ductility
prevent concentration of inelastic deformation at the cold demand. The quasi-static cyclic tests were conducted by
joint between the HPFRC column and the base block, which incrementally increasing the radius of the circular pattern.
was constructed with regular concrete. Such cold joint is The target displacement ductility demand for each cycle in
typical for conventional reinforced concrete bridge columns the displacement history is presented in Table 2. The target
in California. Concentration of deformations at this joint is ductility demand, termed nominal ductility demand μ, was
not desirable because it can lead to a premature sliding shear computed with respect to the yield displacement of the BC
failure at the base of the column, fracture of column longi- column, estimated at 14.0 mm (0.55 in.) (0.86% drift), to
tudinal reinforcement, or both. The cold joint at the base of enable a direct comparison between the BC and HPFRC
HPFRC columns is more vulnerable to damage localiza- specimens. Each post-yield primary cycle (that is, cycle to a
tion than a similar joint in reinforced concrete construction new maximum drift level) was followed by a small displace-
because of the higher flexural strength of the immediately ment cycle with amplitude equal to 1/3 of that of the primary
adjacent HPFRC column sections and the excellent bond cycle to evaluate the column stiffness degradation throughout
between HPFRC and reinforcing bars.9 Thus, the HPFRC the loading history. Similar displacement rates among the
column-foundation interface of the test specimens was different test cycles were applied, not exceeding 25 mm/min
strengthened with dowel reinforcement to force most of the (1.0 in./min). A gravity load equivalent to 0.1fc′Ag of the
inelastic deformations to occur within the HPFRC column. BC column, where Ag is the column gross cross-sectional
area, was applied at the column top through a spreader beam

304 ACI Structural Journal/March-April 2014


Table 2—Nominal ductility level used at each cycle of displacement history, defined with respect to BC
column yield displacement
Cycle Nominal ductility Displacement, mm (in.), (drift, %) Cycle Nominal ductility Displacement, mm (in.), (drift, %)
1 NA 1.0 mm (0.04 in.) (0.07) 10 3 41.9 mm (1.65 in.) (2.6)
2 NA 2.8 mm (0.11 in.) (0.17) 11 1 14.0 mm (0.55 in.) (0.86)
3 NA 5.6 mm (0.22 in.) (0.34) 12 4.5 63.0 mm (2.48 in.) (3.9)
4 1 14.0 mm (0.55 in.) (0.86) 13 1.5 21.1 mm (0.83 in.) (1.3)
5 0.33 4.6 mm (0.18 in.) (0.28) 14 6.25 87.4 mm (3.44 in.) (5.4)
6 1.5 21.1 mm (0.83 in.) (1.3) 15 2 27.9 mm (1.10 in.) (1.7)
7 0.5 7.1 mm (0.28 in.) (0.43) 16 8 112 mm (4.40 in.) (6.9)
8 2 27.9 mm (1.10 in.) (1.7) 17 2 27.9 mm (1.10 in.) (1.7)
9 0.67 9.4 mm (0.37 in.) (0.58) 18 12.5 175 mm (6.88 in.) (10.7)

Fig. 1—Reinforcement detailing of HPFRC specimens. (Note: Dimensions in mm; 1 mm = 0.0394 in.)
and post-tensioned rods tensioned through hydraulic jacks.
This load is typically used to represent the average dead and
service live loads carried by typical column bents of over-
pass bridges in California.

Test setup and instrumentation


The bidirectional quasi-static cyclic testing of the two
HPFRC column specimens was carried out. The lateral and
gravity load test setup is shown in Fig. 3. The lateral load
was applied using two horizontal servo-controlled hydraulic
actuators at an initial angle of 90 degrees with respect to
Fig. 2—Normalized displacement cycle: (a) plan view; and
each other reacting against a rigid frame. The actuators were
(b) displacement history for normalized cycle.
attached to the column top using a steel jacket. The actuator
commands were computed to follow the circular load pattern instrumentation frames located on the sides of the specimen.
considering actual actuator elongations. These absolute displacement measurements were used to
The HPFRC specimen instrumentation scheme comprised obtain the deflected shape of the column and relative deflec-
internal and external instruments. Strain gauges were tions between various column sections, and to monitor the
installed on longitudinal and transverse reinforcement torsion of the column top. Additional linear potentiome-
to trace the strain history at various locations along the ters were attached to instrumentation rods anchored in the
column height and to correlate internal strains to observed column core at various sections to measure the relative
damage, such as bar buckling and fracture. Five levels of displacements necessary for determination of flexural rota-
linear displacement potentiometers were placed on two stiff

ACI Structural Journal/March-April 2014 305


Fig. 3—(a) Plan; and (b) elevation views of experimental setup for HPFRC column specimens.
Table 3—Reinforcing steel mechanical properties
Parameter Continuous longitudinal bar Dowel reinforcement Spiral reinforcement
11 11
Specification ASTM A706 ASTM A706 ASTM A8212
Size No. 4/13M No. 4/13M W3.5
Elastic modulus Es 190.3 GPa (27,600 ksi) 185.5 GPa (26,900 ksi) 198.3 GPa (28,800 ksi)
*
Yield stress Fy 448 MPa (65 ksi) 483 MPa (70 ksi) 638 MPa* (92 ksi)
Ultimate stress Fu 621 MPa (90 ksi) 705 MPa (102 ksi) 709 MPa (102 ksi)
Yield plateau, strain 0.25 to 1.33 — —
*
Yield strength determined using 0.2% offset line parallel to corresponding elastic modulus.

tions, average curvatures, and shear deformations, as well acceptable workability. The column concrete had a slump
as average column axial deformations. Load cells were used of 140 mm (5.5 in.) following the addition of the high-range
to monitor the forces in the two horizontal actuators and the water reducing admixture and before the addition of fibers.
two post-tensioned rods used for simulation of gravity load. Fiber-reinforced concrete 150 x 300 mm (6 x 12 in.) cylin-
ders were cast during the construction of the specimens to
Reinforcing steel assess the development of concrete compressive strength
The mechanical properties of the longitudinal and spiral through ASTM C39/C39M14 tests. In addition, 150 x 150
reinforcement used for the HPFRC columns, obtained x 600 mm (6 x 6 x 24 in.) beams were tested under third-
through ASTM A37010 standard tension tests, are summa- point loading (457 mm [18 in.] span length) following
rized in Table 3. The fracture strain was not determined in ASTM C1609/C1609M15 to assess the flexural performance
these coupon tests because the bars were unloaded before (and indirectly tensile performance) of the HPFRC material
fracture due to laboratory safety regulations. Instead, frac- used. Failure occurred inside the middle third of the spans
ture strain values specified in Caltrans SDC8 were used in for all beam specimens tested. The elastic modulus for the
analytical models of the HPFRC cantilever columns. HPFRC material, Ec,FRC, corresponding to the secant modulus
up to 0.45fc,FRC′, determined from fiber-reinforced concrete
Fiber-reinforced concrete cylinder tests, was also verified through ASTM C1609/
The concrete mixture used for the construction of the C1609M15 beam tests. Average compressive strength test
fiber-reinforced concrete columns and plain concrete foun- results for plain and fiber-reinforced concrete at various days
dation had a specified 28-day strength of 34.5 MPa (5.0 ksi). after casting, obtained directly from ASTM C39/C39M14
This mixture proportion, with a proportion by weight of cylinder tests or from linear interpolation of results at other
0.45:1:2.3:1.86 (water:cement:fine aggregate:coarse aggre- days, are listed in Table 5. Measured cylinder compressive
gate), was adjusted from that used in previous tests6 to incor- strengths at test day are listed in Table 1. Despite having
porate the steel fibers with acceptable workability. Hooked the same specified strength, measured cylinder compres-
steel fibers were added to the concrete mixture at a 1.5% sive strength at test day for the HPFRCs was approximately
volume fraction (120 kg/m3 [200 lb/yd3]) by direct pouring 20% higher than that of the regular concrete used in the
into the truck mixer and mixing for 4 minutes. Mixture BC specimen (Table 1). The elastic modulus, and peak and
quantities for 0.765 m3 (1 yd3) of concrete are listed in residual flexural strength at various deflection levels for the
Table 4. A similar mixture was successfully used in HPFRC fiber-reinforced concrete material determined according to
slab-column connection tests.13 A high-range water reducing ASTM C39/C39M14 and ASTM C1609/C1609M,15 respec-
admixture was added during the mixing process to maintain tively, are also listed in Table 5.

306 ACI Structural Journal/March-April 2014


TEST RESULTS ment level, while the HPFRC columns sustained relatively
The damage progress in the HPFRC and BC specimens minor damage with no spalling despite experiencing rela-
throughout the loading history is shown in Table 6. The tively large flexural cracks.
HPFRC specimens, particularly Specimen S1, exhibited At a nominal ductility level of 6.25 (5.4% drift), the spiral
enhanced damage tolerance compared with the geometri- reinforcement in the S1 column fractured at an approximate
cally identical plain concrete Specimen BC. Both HPFRC height of 150 mm (6 in.) above the column base. This spiral
columns behaved elastically up to a drift ratio of approxi- fracture resulted in longitudinal bar buckling and significant
mately 1.3% (nominal ductility μ = 1.5 based on first yielding concrete cover spalling and core degradation in the subse-
of Specimen BC). The actual ductility of the tested HPFRC quent displacement cycles. All the continuous longitudinal
columns μa is, therefore, two-thirds that of the BC column. bars in Specimen S1 underwent buckling and fractured by
The HPFRC specimens developed relatively similar damage the end of the cycle, corresponding to a nominal ductility
states as the reference BC specimen, but at higher displace- level of 12.5 (10.7% drift). The dowels in Specimen S1
ment levels. For example, damage observed in Specimens did not buckle or fracture. Specimen S2, on the other hand,
S1 and S2 at 3.9% drift ratio (μ = 4.5) was smaller than that experienced longitudinal bar buckling and spiral fracture at
in Specimen BC, as seen in Table 6. Large portions of the two locations, followed by fracture of two continuous longi-
BC column cover had already spalled off at this displace- tudinal reinforcement bars during the cycle at a nominal
ductility demand level of 6.25 (5.4% drift). The BC specimen
Table 4—Fiber-reinforced concrete mixture did not experience spiral fracture, bar buckling, or longitu-
quantities dinal reinforcing bar fracture because the spiral spacing was
Item Value half that provided in the HPFRC specimens, and the column
was only cycled up to a nominal ductility level of 4.5 (3.9%
w/c 0.45
drift) because it was subsequently tested under monotoni-
Coarse aggregate weight (9.5 mm [3/8 in.]
602 kg/m3 (1324 lb/yd3)
cally increased axial load.6 The maximum displacement
maximum size) ductility level attained during the test of Specimens S1 and
Top sand weight 551 kg/m3 (1212 lb/yd3) S2 without substantial loss of gravity load-carrying capacity
Blend sand weight 194 kg/m3 (426 lb/yd3)
was 12.5 and 6.25, respectively (10.7 and 5.4% drift ratio).
Tests of Specimens S1 and S2 were terminated after several
Total fine aggregate weight 745 kg/m3 (1638 lb/yd3) of the longitudinal bars fractured, compromising the safety
Cement 324 kg/m3 (712 lb/yd3) of the test setup.
High-range water-reducing admixture 621 mL/m3 (21 oz/yd3) The length of the plastic deformation zone Lp in Specimen
S1 was approximately 450 mm (18 in.) at the end of the test,
Water 145 kg/m3 (320 lb/yd3)
which is slightly larger than the column diameter. Inelastic
Fibers 88 kg/m3 (194 lb/yd3) deformations in Specimen S1 first developed at approx-
Note: Specified 28-day compressive strength: 34.5 MPa (5.0 ksi); slump: 140 mm
imately 350 mm (14 in.) from the base, and propagated
(5.5 in.); 1 yd3 = 0.765 m3; 1 lb = 0.45 kg. upwards as well as down towards the base of the column.

Table 5—Average values of HPFRC mechanical properties


Parameter 11 days 28 days 49 days 60 days
Regular concrete compres-
sive strength fc′ used for 28.3 MPa* (4.10 ksi) 34.0 MPa (4.93 ksi) 35.0 MPa (5.07 ksi) 35.2 MPa* (5.10 ksi)
BC specimen
Regular concrete compressive
strength fc′ used for HPFRC 33.0 MPa (4.78 ksi) 41.9 MPa (6.07 ksi) 42.1 MPa (6.11 ksi) 41.9 MPa (6.08 ksi)
specimens
HPFRC compressive strength
37.0 MPa (5.37 ksi) 46.7 MPa (6.77 ksi) 47.1 MPa* (6.83 ksi) 47.3 MPa* (6.86 ksi)
fcFRC′
HPFRC elastic modulus Ec-FRC 22.9 GPa* (3323 ksi) 28.9 GPa* (4190 ksi) 29.2 GPa* (4227 ksi) 29.3 GPa* (4246 ksi)
HPFRC first peak flexural
— 5.44 MPa (0.79 ksi) 6.67 MPa (0.97 ksi) 6.37 MPa (0.92 ksi)
strength f1
HPFRC absolute peak flexural
— 6.69 MPa (0.97 ksi) 7.10 MPa (1.03 ksi) 7.58 MPa (1.10 ksi)
strength f
HPFRC residual flexural
strength at L/600 deflection — 6.37 MPa (0.92 ksi) 6.49 MPa (0.94 ksi) 7.29 MPa (1.06 ksi)
fres,L/600
HPFRC residual flexural
strength at L/150 deflection — 4.69 MPa (0.68 ksi) 5.58 MPa (0.81 ksi) 5.38 MPa (0.78 ksi)
fres,L/150
*
Obtained indirectly from linear interpolation of test results performed on other days.

ACI Structural Journal/March-April 2014 307


Table 6—Comparison of damage progress in HPFRC and BC specimens
S1: HPFRC S2: HPFRC BC: Plain concrete
State of column specimens at nominal ductility demand of 1.5 (1.3% drift)

State of column specimens at nominal ductility demand of 2.0 (1.7% drift)

State of column specimens at nominal ductility demand of 3.0 (2.6% drift)

State of column specimens at nominal ductility demand of 4.5 (3.9% drift)

State of column specimens at nominal ductility demand of 6.25 (5.4% drift)

Test terminated.

State of column specimens at nominal ductility demand of 8.0 (6.9% drift)

Test terminated.

State of column specimens at nominal ductility demand of 12.5 (10.7% drift)

308 ACI Structural Journal/March-April 2014


Fig. 4—Force-deformation response of HPFRC and BC specimens: (a) total shear stress versus total drift; and (b) shear stress
ratio versus total drift.
The S2 column, on the other hand, developed a primary and the amount of transverse reinforcement was reduced by
crack at a height of 250 mm (10 in.) above the foundation at half compared with that in the BC specimen, the HPFRC
the cutoff point of the dowel bars, resulting in a concentra- specimens maintained a stable hysteretic behavior governed
tion of deformation and damage in that region with limited by flexure throughout the entire loading history. The degra-
spreading of the plastic deformation zone towards the base dation of lateral resistance with the progression of damage
of the column. The extent of the BC column plastic defor- in the HPFRC specimens, which initiated at a nominal
mation zone was estimated at 300 mm (12 in.), equivalent to displacement ductility of 6.25 (5.4% drift), was governed by
3/4 of the column diameter.6 the loss of flexural capacity (concrete crushing, bar buckling,
and finally bar fracture), and not by shear-related cracking or
Force-displacement response damage. Sliding of the column along the cold joint at the
The force-deformation response of the HPFRC and BC column-foundation interface or significant shear distortions
specimens is shown in Fig. 4. Only the primary cycles corre- of the column plastic hinge zone were not observed.
sponding to the nominal ductility level of 1 and higher are A conventional shear force versus nominal displacement
shown. The deformation axis is expressed in terms of resul- ductility envelope, derived by plotting the peak lateral force
tant drift ratio (displacement of the column top divided by at each nominal displacement ductility level imposed, is
the distance between the top of the column base to the actu- shown in Fig. 5(a). This plot was computed using the peak
ator axes). The force axis is expressed in terms of resultant resultant force of the specimens in the x- and y-directions,
shear stress (Fig. 4(a)) and shear stress ratio (Fig. 4(b)). The to reconcile the differences in the response produced by the
resultant shear stress was calculated as the resultant shear circular load pattern. Both HPFRC specimens remained in
force applied by the two actuators divided by the effective the elastic response range up to a nominal ductility demand
area in shear, defined by Caltrans SDC8 as 0.8 times the gross of 1.5 (based on first yielding of Specimen BC). The peak
area of the column. The shear stress ratio was obtained by applied force remained generally constant between nominal
dividing the total shear force in the column by the assumed ductility levels of 1 and 4.5 for the BC column, and between
nominal shear strength provided by the transverse steel rein- 1.5 and 6.25 for the HPFRC columns. The S1 column was
forcement at yielding Vs, calculated according to Caltrans slightly stronger than the S2 column.
SDC8 as follows The secant lateral stiffness of the specimens, computed
using the resultant peak force and corresponding displace-
Av f yh D ′ ment value attained during each new primary cycle, is
Vs = (1) shown in Fig. 5(b). The cracked elastic stiffness of the two
s
HPFRC specimens was comparable, and higher than that of
Specimen BC. However, the stiffness degradation rate for
where Av = (π/2)Ab is the area of shear reinforcement,
Specimen BC was somewhat lower than that for the HPFRC
Ab is the area of the spiral reinforcement, fyh is the corre-
specimens up to nominal displacement ductility of 4.5. This
sponding yield strength, D′ is the cross-sectional dimension
is likely due to the use of twice as much transverse reinforce-
of confined concrete core measured between the centerline
ment (and tighter spiral spacing) in Specimen BC.
of the spiral, and s is the spiral pitch.
Figure 4 shows a three-fold increase in shear stress ratio
Column plastic hinge deformations
demand in the HPFRC specimens compared with the BC
Measurements of cross section rotation relative to the
specimen due to: 1) half as much transverse reinforcement in
column base about the x- and y-axes were obtained at four
the HPRFC columns than in the BC column; and 2) dowels
levels along the height of the column (0.375Dcol, 0.75Dcol,
in the plastic hinge region of the HPRFC specimens that
1.125Dcol, and 4.5Dcol above the column base). From these
added flexural resistance and moved up the plastic hinge
rotations, average curvatures for each segment between two
region, increasing the shear demand at flexural yielding.
adjacent rotation measurements were calculated to eval-
From Fig. 4, it can be seen that even though the shear
uate the spread of inelastic deformations along the column
demand on the HPFRC specimens increased significantly

ACI Structural Journal/March-April 2014 309


Fig. 5—(a) Shear force-displacement envelope; and (b) lateral stiffness degradation versus nominal displacement ductility (or
drift) demand.

Fig. 6—Curvature profiles from experiments for: (a) S1; (b) S2; and (c) BC.
height. Profiles of these curvatures for all three specimens HPFRC plastic hinges. It should be noted, however, that
are shown in Fig. 6. increased curvature demands were imposed on the HPRFC
The significant difference in the length of the plastic specimens for a given displacement level, compared with
deformation zones among these three specimens, discussed those imposed on the BC specimen, due to the upward shift
previously and evident in Fig. 6, shows that it is possible of the plastic deformation region.
to design and detail the longitudinal reinforcement of an
HPFRC column to achieve a highly desirable spreading of Bond stress
plastic deformation. In particular, the addition of dowel rein- In both HPFRC specimens, the bonded region of dowel
forcement elevated the center of the plastic hinge zone from reinforcement started at 250 mm (10 in.) above the founda-
the column base, thus providing more space for its spreading. tion. Strain gauge measurements along the dowel reinforce-
Debonding of the dowel reinforcement (thus avoiding termi- ment recorded peak strain values exceeding the steel yield
nation of dowels within the plastic hinge) in the column of strain at a height of 100 mm (4 in.) above the foundation or
Specimen S1 allowed for very effective spreading of bar 150 mm (6 in.) from the bar termination point, resulting in
yielding along the column height and formation of several a length Ld as small as 150 mm (6 in.) or 12 bar diameters
flexural cracks in the plastic hinge region. The less successful required to develop the yield strength of the No. 4 dowel
detail used in Specimen S2, on the other hand, shows that bars. The peak average bond stress up to first yielding of
there are significant unexplored opportunities to develop the reinforcement in the HPFRC specimens was determined
improved designs to ensure adequate spread of yielding in based on force equilibrium between the resultant force from

310 ACI Structural Journal/March-April 2014


Fig. 7—Curvature profiles for calibrated plastic hinge model of Column S1: (a) formulation; and (b) results for different
displacement ductility demands of HPFRC columns.
an average uniform bond stress ub acting on the surface of flexural yielding, a uniform curvature over the plastic hinge
the dowel reinforcement along a length measured from the length is assumed, dependent on the actual ductility level of
bar termination point to the section at which first yielding the HPFRC column μa. Curvature over the length hLP is also
occurred, and the yield strength of the bar. It is important assumed constant, as shown in Fig. 7, and equal to 1/4 of the
to note that dowel bar yielding occurred after substan- curvature over Lp. Additional information about the material
tial cracking occurred around the cover of the HPFRC models and parameters used for calibration of the HPFRC
specimens. The resulting peak average bond stress ub was plastic hinge model can be found elsewhere.6,7 A difference
10.2 MPa (1.5 ksi), which can be rewritten in terms of the of less than 10% was obtained in the computation of lateral
unconfined HPFRC compressive strength results obtained displacements, u, when using the experimentally (Fig. 6(a))
from cylinder tests, fc′,FRC = 47.3 MPa (6.86 ksi). The degra- and analytically (Fig. 7) obtained curvatures. In the ideal-
dation of bond stress with increasing bar inelastic strains ized curvature profile shown in Fig. 7(a), the term HTot is the
could not be evaluated due to lack of data. Thus, the peak height of the column measured from the base block top to
bond stress for reinforcing bar strains not exceeding the yield the actuator centerline.
strain was approximately ub,max = 1.5 fc′, FRC , MPa (ub,max =
18 fc′, FRC , psi). This value is significantly higher than bond SUMMARY AND CONCLUSIONS
strengths reported in the literature for regular concrete16 of Experimental findings obtained from tests of strain-hard-
1.0√fc′, MPa (12√fc′, psi) and 0.5√fc′, MPa (6√fc′, psi) for ening, high-performance fiber-reinforced concrete (HRFRC)
deformed bars at slip values smaller and larger than the slip bridge columns under highly demanding bidirectional cyclic
measured at bar yield strain, respectively. These measure- displacements are presented. Two approximately 1/4-scale
ments thus indicate that reinforcement development lengths circular cantilever HPFRC column specimens (S1 and S2)
in HPFRC columns could be shorter than those in conven- were tested, and the results were compared with those of a
tional concrete columns. Similar conclusions were estab- geometrically identical baseline regular concrete specimen,
lished in recent studies.9 Conservative development lengths denoted as Specimen BC. The HPFRC column specimens
for conventional concrete, however, should be used for were constructed using a ready mix concrete with a 1.5%
HPFRC until more test data on the subject become available. volume fraction of high-strength hooked steel fibers. The
plastic hinge region of the HPFRC specimens was detailed
CALIBRATED HPFRC COLUMN MODELS using dowels in combination with bar debonding, either of
An idealized curvature profile for columns with rein- the dowel ends or of the main longitudinal reinforcement at
forcement detailing similar to that used in Specimen S1 is the end of the dowels to prevent concentration of damage
shown in Fig. 7. Based on the curvature distribution shown at the column base and increase the spread of yielding. The
in Fig. 6(a), the length of the plastic hinge zone Lp was set at transverse reinforcement spacing in the HPFRC specimens
400 mm (16 in.), equal to the column diameter. The middle was twice that of the BC specimen.
of the plastic hinge was centered at approximately the middle The HPFRC specimen with long dowels, Specimen S1,
of the debonded region of the column dowel bars, by setting developed an extended plastic deformation zone with an
the distance between the column base and the bottom of the approximate length of one column diameter, improved
plastic hinge zone, hLP, to 150 mm (6 in.). The yield curva- ductile behavior, high-damage tolerance, and high energy
tures φy,Top and φy,Bottom were defined at a height of 560 and dissipation compared with the regular concrete Specimen
150 mm (22 and 6 in.) above the column base (top and bottom BC. This specimen was cycled up to a 10.7% drift ratio
ends of the plastic hinge zone), respectively. These yield while sustaining the applied gravity load. The HPFRC spec-
curvatures were computed using moment-curvature analyses imen with short dowels and debonding of main longitudinal
of the corresponding cross sections with different longitu- reinforcement, Specimen S2, exhibited less deformation
dinal reinforcement details. For displacements beyond first capacity compared with Specimen S1. The reason for such

ACI Structural Journal/March-April 2014 311


behavior was the development of a single major crack at the Gustavo J. Parra-Montesinos, FACI, is the C. K. Wang Professor of
Structural Engineering at the University of Wisconsin-Madison, Madison,
cutoff point of the dowel bars, resulting in a concentration of WI. He is Chair of ACI Committee 335, Composite and Hybrid Structures;
rotation and damage at that location and limited spreading of and a member of ACI Committees 318, Structural Concrete Building Code,
the plastic hinge zone towards the base of the column. and 544, Fiber-Reinforced Concrete, and Joint ACI-ASCE Committee 352,
Joints and Connections in Monolithic Concrete Structures. His research
Even though the HPFRC specimens were subjected to interests include seismic behavior and design of reinforced concrete,
larger shear force demands and the amount of transverse fiber-reinforced concrete, and hybrid steel-concrete structures.
reinforcement was half that of the conventional Specimen
BC, the HPFRC specimens exhibited a stable hysteretic REFERENCES
behavior governed by flexure up to large drift demands with 1. Parra-Montesinos, G. J., “High-Performance Fiber-Reinforced
Cement Composites: An Alternative for Seismic Design of Structures,” ACI
negligible shear-related damage. Such desirable structural Structural Journal, V. 102, No. 5, Sept.-Oct. 2005, pp. 668-675.
response of the tested HPFRC columns indicates that HPFRC 2. Billington, S. L, and Yoon, J. K., “Cyclic Response of Unbonded
has great potential for use in flexural elements subjected Post-Tensioned Precast Columns with Ductile Fiber-Reinforced Concrete,”
Journal of Bridge Engineering, ASCE, V. 9, No. 4, July-Aug. 2004,
to large shear reversals. The results of these unique tests pp. 353-363.
demonstrate several other advantages of HPFRC and call 3. Harajli, M. H., and Rteil, A. M., “Effect of Confinement Using
for additional research on the design and detailing of plastic Fiber-Reinforced Polymer or Fiber-Reinforced Concrete on Seismic
Performance of Gravity Load-Designed Columns,” ACI Structural Journal,
hinge regions in HPFRC flexural members to best achieve V. 101, No. 1, Jan.-Feb. 2004, pp. 47-56.
improved seismic performance and reduced post-earthquake 4. Canbolat, B. A.; Parra-Montesinos, G. J.; and Wight, J. K., “Experi-
repair costs of bridge structures. Furthermore, the use of mental Study on Seismic Behavior of High-Performance Fiber-Reinforced
Cement Composite Coupling Beams,” ACI Structural Journal, V. 102,
HPFRC is expected to simplify the construction of critical No. 1, Jan.-Feb. 2005, pp. 159-166.
regions in bridges by allowing for a substantial increase 5. Parra-Montesinos, G. J., and Chompreda, P., “Deformation Capacity
in transverse reinforcement spacing compared with that and Shear Strength of Fiber-Reinforced Cement Composite Flexural
Members Subjected to Displacement Reversals,” Journal of Structural
required in regular concrete construction without compro- Engineering, ASCE, V. 133, No. 3, 2007, pp. 421-431.
mising seismic performance. 6. Terzic, V., and Stojadinovic, B., “Post-Earthquake Traffic Capacity of
While the additional shear strength provided by the steel Modern Bridges in California,” PEER Report 2010/103, Pacific Earthquake
Engineering Research Center, Berkeley, CA, 2010, 218 pp.
fibers is helpful in ensuring a flexural-dominated behavior, 7. Aviram, A.; Stojadinovic, B.; Parra-Montesinos, G. J.; and Mackie,
the adverse consequences of increased hoop spacing, partic- K. R., “Structural Response and Cost Characterization of Bridge Construc-
ularly reduction of lateral support of longitudinal rein- tion Using Seismic Performance Enhancement Strategies,” PEER Report
2010/01, Pacific Earthquake Engineering Research Center, Berkeley, CA,
forcing bars, should be carefully considered during design 2009, 363 pp.
of HPFRC elements. The efficiency of HPFRC cover to 8. Caltrans, “Seismic Design Criteria 1.3,” Report, California Depart-
provide support to the longitudinal bars is as yet unknown, ment of Transportation, Sacramento, CA, 2004, 101 pp.
9. Chao, S.-H.; Naaman, A. E.; and Parra-Montesinos, G. J., “Bond
and should not be relied upon until results from research on Behavior of Reinforcing Bars in Tensile Strain-Hardening Fiber Reinforced
this topic become available. Cement Composites,” ACI Structural Journal, V. 106, No. 6, Nov.-Dec.
2009, pp. 897-906.
10. ASTM A370, “Standard Test Methods and Definitions for Mechan-
AUTHOR BIOS ical Testing of Steel Products,” ASTM International, West Conshohocken,
ACI member Ady Aviram is a Structural Engineer with Simpson Gumpertz
PA, 2012, 48 pp.
& Heger, Inc., in San Francisco, CA. She received her BS in civil engi-
11. ASTM A706/A706M, “Standard Specification for Low-Alloy Steel
neering from the University of Costa Rica, San Pedro, Costa Rica, and her
Deformed and Plain Bars for Concrete Reinforcement,” ASTM Interna-
MEng and PhD in structural engineering from the University of California,
tional, West Conshohocken, PA, 2009, 6 pp.
Berkeley, Berkeley, CA. Her research interests include performance-based
12. ASTM A82, “Standard Specification for Steel Wire, Plain, for
earthquake-resistant design of steel and reinforced-concrete structures,
Concrete Reinforcement,” ASTM International, West Conshohocken, PA,
base-plate connections, bridge modeling and analysis, structural reliability,
2007, 4 pp.
fiber-reinforced concrete, and blast-resistant design of wall systems.
13. Cheng, M.-Y.; Parra-Montesinos, G. J.; and Shield, C. K., “Shear
Strength and Drift Capacity of Fiber Reinforced Concrete Slab-Column
Bozidar Stojadinovic, FACI, is a Professor and Chair of structural dynamics Connections Subjected to Bi-Axial Displacements,” Journal of Structural
and earthquake engineering at the Swiss Federal Institute of Technology Engineering, ASCE, V. 136, No. 9, 2010, pp. 1078-1088.
(ETH), Zürich, Switzerland. He received his Dipl. Ing. from the University 14. ASTM C39/C39M, “Standard Test Method for Compressive
of Belgrade, Belgrade, Serbia; his MS from Carnegie Mellon University, Strength of Cylindrical Concrete Specimens,” ASTM International, West
Pittsburgh, PA; and his PhD from the University of California, Berkeley. He Conshohocken, PA, 2012, 7 pp.
is a member of ACI Committees 335, Composite and Hybrid Structures; 341, 15. ASTM C1609, “Standard Test Method for Flexural Performance
Earthquake-Resistant Concrete Bridges; 349, Concrete Nuclear Structures; of Fiber-Reinforced Concrete (Using Beam With Third-Point Loading),”
and 374, Performance-Based Seismic Design of Concrete Buildings. His ASTM International, West Conshohocken, PA, 2012, 9 pp.
research interests include probabilistic performance-based seismic design 16. Eligehausen, R.; Popov, E. P.; and Bertero, V. V., “Local Bond
of composite and reinforced concrete structures. Stress-Slip Relationships of Deformed Bars under Generalized Exci-
tations,” Report 82/23, Earthquake Engineering Research Center, Univer-
sity of California, Berkeley, Berkeley, CA, 1983, 169 pp.

312 ACI Structural Journal/March-April 2014


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 111-S27

Analysis of Early-Age Thermal and Shrinkage Stresses in


Reinforced Concrete Walls
by Barbara Klemczak and Agnieszka Knoppik-Wróbel
The issues related to thermal and shrinkage stresses arising in the phase of temperature increase, tensile stresses originate
reinforced concrete walls at the stage of their construction are in the surface layers of the element, and compressive stresses
discussed. The cause of these stresses is inhomogeneous volume are observed inside the element. An inversion of stress body
changes associated with temperature rise caused by the exothermic occurs during the cooling phase: inside, tensile stresses
hydration process of cement, as well as moisture exchange with
are observed; in the surface layers, compressive stresses
the environment. The induced thermal-shrinkage stresses can
are observed. Considerable self-induced stresses can be
reach significant levels, and in some cases, cracks appear in
structural elements. expected, for example, within thick foundation slabs, thick
The article presents the results of a numerical analysis of a rein- walls, dams, and in each element with interior temperatures
forced concrete wall cast against an old set foundation, subjected considerably greater than surface temperatures.
to early-age thermal and shrinkage deformations. Development of A concrete element can also be externally restrained. For
stresses and character of cracks is briefly described. The presented example, such a restraint exists along the contact surface of
analysis focuses on evaluation of the contribution of self-induced mature concrete against which a new concrete element has
and restraint stresses to the total stresses induced in the wall. The been cast. Because of the limited possibility of deforma-
contribution and development of thermal and shrinkage stresses is tions of the structure, restraint stresses occur in that case.
investigated. Walls with different dimensions are considered. The restraint stresses are often observed in medium-thick
Keywords: cracking; early-age concrete; reinforced concrete wall; thermal-
elements, such as a wall cast against an old set concrete.
shrinkage stresses. It should also be noted that the stresses resulting from an
external restraint of a structure add to the effects of an
INTRODUCTION internal restraint.
The cause of thermal and shrinkage stresses arising The problem of thermal-shrinkage stresses—and conse-
in early-age concrete are the volume changes due to the quently, in some cases, cracking of early-age concrete
temperature and moisture variations during the hardening structures—is well known in massive concrete elements.1-3
process. The variations of concrete temperature during Nevertheless, high thermal-shrinkage stresses and cracks are
curing are the result of the exothermic nature of the chemical also observed in newly constructed medium-thick concrete
reaction between cement and water. In structural elements elements, such as reinforced concrete (RC) walls cast against
with thin sections, the generated heat dissipates quickly, and an old set foundation (tank walls or abutments4,5) or walls cast
causes no problem. In thicker sections, the internal tempera- in stages (massive container walls6). These cracks, which are
ture can reach a significant level. Furthermore, due to the particularly deep or thorough cracks, may adversely affect
poor thermal conductivity of concrete, high temperature the serviceability, lifespan, or even bearing capacity of a
gradients may occur between the interior and the surface of concrete structure. Cracking in tank walls endangers their
thick structural elements. Concrete curing is also accompa- tightness.5 The problem of thermal-moisture cracking is also
nied by moisture exchange with the environment in condi- dangerous in the case of nuclear containments executed in
tions of variable temperatures. The loss of water through stages, where a previously cast layer of mature concrete
evaporation at the surface of the element results in shrinkage, restrains younger concrete layers.6 The formation of cracks
which is classified as an external drying shrinkage. There promotes the leakage of radioactive elements into the envi-
is also internal drying resulting from the reduction in mate- ronment during the service life of the containment. Early-age
rial volume as water is consumed by hydration, which is thermal-shrinkage cracking may even influence thin-walled
classified as autogenous shrinkage. Additionally, the chem- structures, such as shell roof covers, which reduces their
ical shrinkage is also distinguished, which occurs because capacity, and may lead to structural collapse as a result of
the volume of hydration products is less than the original unforeseen overpressure.7
volume of cement and water.
The volume changes due to the temperature and moisture RESEARCH SIGNIFICANCE
variations have consequences in arising stresses in a concrete Considerable early-age thermal-shrinkage stresses and
element. Two natures of these stresses can be distinguished: consequent cracks at the construction stage are frequently
self-induced stresses and restraint stresses. ACI Structural Journal, V. 111, No. 2, March-April 2014.
The self-induced stresses are related to internal restraints MS No. S-2012-096.R2, doi:10.14359.51686523, was received December 14,
2012, and reviewed under Institute publication policies. Copyright © 2014, American
of the structure, resulting from nonuniform volume changes Concrete Institute. All rights reserved, including the making of copies unless
permission is obtained from the copyright proprietors. Pertinent discussion including
in a cross section. In internally restrained elements, during author’s closure, if any, will be published ten months from this journal’s date if the
discussion is received within four months of the paper’s print publication.

ACI Structural Journal/March-April 2014 313


Fig. 1—(a) Temperature; (b) thermal stresses; (c) moisture content; and (d) shrinkage stress development in time for externally
restrained concrete wall.10
observed in RC walls. Correct prediction and counteracting with the height of the wall. Nevertheless, the main character
negative effects of these stresses is of great importance to of temperature, moisture, and stress development presented
ensure the desired service life and function of structures. It in Fig. 1 is kept in most areas of the wall. The detailed
is particularly important in structures such as tanks, which analysis of the temperature, moisture, and stress distribu-
require a solid concrete that prevents water leakage. This tion in the walls with different dimensions are presented in
article studies distribution of the discussed stresses in RC following sections of the article.
walls with different dimensions. The contribution of self-in- Shrinkage stresses resulting from moisture migration
duced and restrained stresses to the total induced stresses is reach relatively low values, so the character of total ther-
also investigated. mal-shrinkage stresses is analogical to thermal stresses;
shrinkage stresses increase total tensile stress values in the
DEVELOPMENT OF THERMAL AND SHRINKAGE second phase. Figure 2 presents total thermal-shrinkage
STRESSES stress maps with resulting deformations in the right half
Two main phases can be distinguished when observing of the 70 cm (27.56 in.) thick wall in a heating phase after
a temperature change in time during the concrete curing 1.3 days (Fig. 2(a)) and a cooling phase after 18.3 days
process (Fig. 1(a)): a phase of concrete temperature increase (Fig. 2(b)).
(self-heating), and a cooling phase of the element down to The described tensile stresses occurring in the cooling
the temperature of the surrounding air. In the first phase, stage often reach considerable values, and can lead to
the wall extends being opposed by the foundation, which cracking of the wall. A typical pattern of cracking due to an
results in the formation of compressive stresses (Fig. 1(b)), edge restraint of a wall is shown in Fig. 3, assuming that the
usually within the first 1 to 3 days. As soon as the maximum base is rigid. Without restraint, the section would contract
self-heating temperature is reached, the wall begins to cool along the line of the base; thus, with a restraint, a horizontal
down, which takes another few days, restrained by a cooled force develops along the construction joint. This leads to
foundation. This leads to development of tensile stresses in vertical cracking at the midspan area, but splayed cracking
the wall (Fig. 1(b)). In case of moisture migration, a mono- towards the ends of the section where a vertical tensile force
tonic moisture removal from the wall is observed (Fig. 1(c)). is required to balance the tendency of the horizontal force
The resulting stresses are tensile stresses in the whole curing to warp the wall. In addition, a horizontal crack may occur
process (Fig. 1(d)). at the construction joint at the ends of the walls due to this
It should be noted that the graphs in Fig. 1 are shown only warping restraint.
for illustration of the phenomena that arise in the discussed Generally, it is thought that a basic cracking pattern is
RC walls. In fact, the values of generated temperature and independent of the amount of reinforcement provided.4,5,8
the loss of moisture will be different in each point of the wall When sufficient reinforcement is provided, the widths of
according to the temperature and moisture exchange with the cracks are controlled, although secondary cracks may be
environment. Similarly, the values of generated stresses vary induced. The extent and size of cracking will then depend
in particular areas of the wall due to the different thermal on the amount and distribution of reinforcement provided.
and shrinkage strains, as well as due to the different level of The cracks can reach 1/3, 1/2, or even 2/3 of the height of
the restraint of the wall in the foundation, which is changing the wall, depending on the length and height of the wall, and

314 ACI Structural Journal/March-April 2014


Fig. 2—Distribution of thermal-shrinkage stresses in right half of reinforced concrete wall in: (a) heating phase; and (b)
cooling phase.
• Conditions during concreting and curing of concrete,
such as the initial temperature of concrete, type of form-
work, and use of insulation or pipe cooling;
• Technology of concreting, such as segmental concreting;
• Environmental conditions, such as ambient tempera-
ture, temperature of neighboring elements, wind, and
humidity; and
• Dimensions of wall, especially the width of the wall.
Fig. 3—Image, spacing, and height of cracks in reinforced Furthermore, deformations originating from the material
concrete wall.4 from which a wall is made and occurring in the erection
stage are essential loads. The problems arise from the mate-
are usually spaced every 1.5 to 3.0 m (4.9 to 9.8 ft).4 The rial itself as concrete is subjected to transformations caused
cracks start at the wall-foundation interface and widen up to by cement hydration and its mechanical properties develop
the wk,max value, and then decrease in width. The maximum as its maturity progresses. Therefore, first the nonlinear and
width of the crack wk,max is approximately 0.3 to 0.5 mm non-stationary temperature and moisture fields and corre-
(12 to 20 mil) (in walls with a low horizontal reinforcement sponding strains should be determined in the analyzed wall
ratio), and is localized at some level above the construc- considering the real technological and material conditions.
tion joint (at 1/3 of the height of the crack, according to The stresses are calculated in the second step of the analysis.
Reference 4; at a height equal to the thickness of the wall,
according to References 9 and 10). Another interesting prop- METHOD OF ANALYSIS
erty of the cracks is their distribution: the greatest height of The applied original numerical model can be classified as a
the crack can be observed in the middle of the wall length, phenomenological model. The influence of mechanical fields
and it declines towards free edges of the wall or towards the on the temperature and moisture fields was neglected, but
expansion joints, as presented in Fig. 3. thermal-moisture fields were modeled using coupled equa-
tions of thermodiffusion. Therefore, the complex analysis of
PREDICTION OF THERMAL AND SHRINKAGE a structure consists of three steps. The first step is related to
STRESSES determination of temperature and moisture development. In
Prediction and control of thermal and shrinkage stresses, the second step, thermal-shrinkage strains are calculated, and
along with possible cracking in early-age RC walls, is a these results are used as an input for computation of stress in
complicated problem due to the complex nature of inter- the last step. With respect to the engineering application of
acting phenomena and a large number of contributing the theoretical model, computer codes were also developed.
factors.11-14 One of the important factors is the geometry of Details of the model are given in Appendix A,* and a full
the RC wall (length, height, and the length-height ratio), as description of the model and computer programs, TEMWIL
well as a degree of its restraint in a foundation. Addition- and MAFEM_VEVP, can be found in References 15 through
ally, the crucial factor is the temperature development in 17. Some results of the validation of the model are presented
the concrete member. The complex variables that affect the in Appendix B,* where the cracking image obtained in the
rate of temperature rise, the maximum temperature, and the numerical analysis is compared with the cracking observed
temperature gradients over sections of the wall are: in the real tank walls. For presentation of results, an open-
• Thermal properties of early-age concrete, such as the source application PARAVIEW was adopted.
rate of heat evolution, the total amount of heat, specific
heat, and thermal conductivity, are strongly dependent
*
The Appendix is available at www.concrete.org/publications in PDF format,
on the amount and properties of concrete components,
appended to the online version of the published paper. It is also available in hard
especially the amount and type of cement; copy from ACI headquarters for a fee equal to the cost of reproduction plus handling
at the time of the request.

ACI Structural Journal/March-April 2014 315


Table 1—Thermal and moisture coefficients
Thermal fields
Coefficient of thermal conductivity λ, W/mK (Btu/s·ft°F) 2.52 (4.04 × 10–4)
Specific heat cb, kJ/kgK (Btu/lb°F) 0.95 (0.227)
Density of concrete ρ, kg/m (lb/ft )
3 3
2400 (149.80)
Coefficient of thermal diffusion αTT, m /s (ft /s)
2 2
11.1 × 10–7 (1.19 × 10–5)
Coefficient representing influence of
αTW, m2K/s (ft2·°F/s) 9.375 × 10–5 (1.81 × 10–3)
moisture concentration on heat transfer
6.00 (29.41 × 10–5) surface without protection, without
considering wind
Thermal transfer coefficient αp, W/m2K (Btu/ft2s°F) 3.58 (17.56 × 10–5) surface with plywood
5.80 (28.45 × 10–5) surface with foil
0.81 (3.97 × 10–5) bottom surface: soil

According to equation Q (T , t ) = Q∞ e 
 − ate−0 ,5 
Heat of hydration* 
Q∞ = kJ/kg (218.90 Btu/lb); a = 513.62te–0.17

Moisture fields
Coefficient of water-cement
K, m3/J (ft3/Btu) 0.3 × 10–9 (1.12 × 10–5)
proportionality
Coefficient of moisture diffusion αWW, m2/s (ft2/s) 0.6 × 10–9 (6.46 × 10–9)
Thermal coefficient of moisture diffusion αWT, m2/sK (ft2/s°F) 2 × 10–11 (1.20 × 10–10)
2.78 × 10–8 (91.21 × 10–9) surface without protection
0.18 × 10–8 (5.90 × 10–9) surface with plywood
Moisture transfer coefficient βp, m/s (ft/s)
0.10 × 10–8 (3.28 × 10–9) surface with foil
0.12 × 10–8 (3.93 × 10–9) bottom surface: soil
*
Approximation made on basis of experimental results of heat of hydration.

at the free edges of the wall and within the contact surface
between the wall and the foundation. A final geometry of the
wall with a mesh of finite elements for one exemplary wall
is presented in Fig. 4.
It was assumed that the analyzed wall was made of the
following concrete mixture: cement CEM I 42.5R 375 kg/m3
(23.41 lb/ft3), water 170 L/m3 (10.61 lb/ft3), and aggregate
(granite) 1868 kg/m3 (116.60 lb/ft3). Thermal and moisture
coefficients necessary for calculations were set in Table 1.
Fig. 4—Dimensions of analyzed walls with finite element
The development of mechanical properties in time was
mesh.
assumed according to CEB-FIP MC90.18 The final values
ANALYSIS OF STRESSES: RESULTS AND for 28-day concrete were assumed as follows: compres-
DISCUSSION sive strength fcm = 35 MPa (5.08 ksi), tensile strength fctm =
3.0 MPa (0.44 ksi), and modulus of elasticity Ecm = 32.0 GPa
Assumption of geometrical, material, and (4.64 Mpsi). It was also assumed that the foundation was
technological data erected earlier and had hardened, so the material properties
The wall of 4 m (13.12 ft) height was analyzed for the were taken as for 28-day concrete, with the same final values
length of 10, 15, or 20 m (32.81, 49.21, or 65.62 ft) and as the wall. Environmental and technological conditions
two thicknesses: 40 and 70 cm (15.75 and 27.56 in.). The were taken as: ambient temperature 20°C (68°F), initial
length-height ratios of these walls were, respectively, 2.5, temperature of fresh concrete mixture 20°C (68°F), wooden
3.75, and 5. Six examples of combinations of these cases formwork of 1.8 cm (0.71 in) plywood on the side surfaces,
were considered. The analyzed walls were supported on a and foil protection of the top surface. It was also assumed
4 m (13.12 ft) wide and 70 cm (27.56 in.) deep strip foun- that formwork was removed 28 days after concrete casting.
dation of the same length. The wall and the foundation were
assumed to be reinforced with a near-surface reinforcing Thermal and shrinkage stresses
net of ø16 bars (0.63 in.). The wall was reinforced at both First, the temperature and moisture development in
surfaces with horizontal spacing of 20 cm (7.87 in.) and time was determined. Figure 5 presents a juxtaposition of
vertical spacing of 15 cm (5.91 in.). The foundation was temperature and moisture content development diagrams for
reinforced with 20 x 20 cm (7.87 x 7.87 in.) spacing at the two areas in the walls (Fig. 4) with different dimensions.
top and bottom surface. Due to a double symmetry of the Although the character of both temperature and moisture
wall, the model for finite element analysis was created for content are independent of the dimensions of the wall, their
1/4 of the walls. A uniform mesh was prepared and densified magnitudes depend directly on these dimensions. Only the

316 ACI Structural Journal/March-April 2014


thickness, not the length, of the wall is influenced by the
temperature and moisture content development. The greatest
temperatures are reached in the thicker walls, while the
moisture removal rates are almost the same; only slightly
greater loss of moisture was observed in thinner walls.
For the known thermal-moisture fields and strains, the
stress state can be determined. The following cases were
analyzed, and the results of stress distribution at the height of
the wall in its internal midspan cross section were presented:
1. Thermal stresses with assumed uniform distribution of
temperature in the wall (development of temperature in time
was only considered) (Fig. 6(a) and (c));
2. Shrinkage stresses with assumed uniform moisture
content distribution in the wall (moisture content change in
time was only considered) (Fig. 6(b) and (d));
3. Thermal stresses with assumed real distribution of
temperature in the wall (development of temperature in time
was also considered) (Fig. 7(a), (c), and (e));
4. Shrinkage stresses with assumed real distribution of
moisture content in the wall (moisture content change in
time was also considered) (Fig. 7(b), (d), and (f)); and
5. Coupled thermal and shrinkage stresses with assumed
real distribution of both temperature and moisture content
(Fig. 8(a)).
The diagrams of nonuniform temperature distribution at
the height of the wall (Fig. 7(a)) are presented at the moment
the maximum hardening temperature is reached; in case of
uniform distribution of temperature, the values of tempera-
ture from the interior of the wall are applied the entire wall
(Fig. 6(a)). Similarly, the moisture distribution is presented at
the moment the maximum hardening temperature is reached.
Figure 6 presents distribution of thermal (Fig. 6(c)) and
shrinkage (Fig. 6(d)) stresses in the midspan cross section
of the interior of the wall under the assumption of uniform
temperature and moisture content distribution in the wall.
Such an assumption is very common in the analysis of
medium-thick externally restrained structures, especially
when analytic methods of thermal-shrinkage stresses deter-
mination are used, but also in numerical analyses in which
a model is reduced to a two-dimensional problem. It is
believed that such an approach provides a good approxi-
mation because both the temperature and moisture content
differences within the body of the wall are relatively small.
The resultant stress distribution at the height of the section
is approximately linear with the maximum values of stresses
at the joint between the wall and the foundation. On such
a simple example, it can be observed that both the length
and the thickness of the wall influence the resulting stresses.
The thickness of the wall determines the maximum value
of stress; it should be noted that greater values of thermal
stresses occur in thicker walls, which results from higher
exerted temperatures; shrinkage stresses are greater in
thinner walls, which is caused by a higher rate of water
Fig. 5—Temperature distribution: (a) in interior and (b) on
removal. The length of the wall (linear restraint) determines
surface of wall; and moisture content development: (c) in
the distribution of stress at the height of the wall. For the
interior and (d) on surface of wall.
analyzed walls characterized by length-height ratio (L/H) ≥
2.5, tensile stresses occur at the whole height of the wall; in Nevertheless, heat and moisture are transported within the
high walls—that is, the walls with L/H < 2.5—compression element and to the surrounding environment in the process
of top areas may be expected. of concrete curing. Therefore, the values of temperatures

ACI Structural Journal/March-April 2014 317


Fig. 6—Case of uniform distribution of: (a) temperature; (b) moisture; (c) thermal; and (d) shrinkage stress distribution at
height of wall in midspan cross section after 18.3 days.
and moisture content vary in different zones of the wall, as the wall. The character of total thermal-shrinkage stresses
do the resulting stresses. Figure 7 shows thermal (Fig. 7(c)) results mainly from the character of their thermal compo-
and shrinkage (Fig. 7(d)) stress distribution at the height of nent (Fig. 7(b) versus Fig. 8(a)); shrinkage stresses only add
the wall in its interior taking into account the real (nonuni- to the final value. Thus, the aforementioned conclusions
form) temperature and moisture content distribution. It is remain valid. The location of the maximum stresses varied;
also noted that temperature and moisture content difference generally, it was the closest to the construction joint for the
at the thickness of the wall leads to stress diversification in thinnest and the shortest walls (0.4 m = 1.31 ft), and was
the internal and near-surface areas (Fig. 7(e) and (f)). It can elevated as the thickness and the length of the wall increased
be observed that the length of the wall influences the occur- (even up to 1.2 m [3.94 ft]). Thus, the results comply with
ring stresses as much as its thickness in such a way that the observations in Reference 4, while the observations from
length determines the character of stress distribution, while References 9 and 10 seem more accurate for high walls
the thickness determines the maximum values of stresses. (walls with a low L/H ratio). In all analyzed cases, because
It should be emphasized that considering real distribution the formwork was detained for the whole process, greater
of temperature and moisture the maximum stresses, and total stresses were observed in the interior of the wall, which
consequently the highest cracking risk, is observed at some explains the occurrence of first cracks in the interior of
distance above the joint, which complies with observations the wall.5
in References 4, 9, and 10. It should be noted that Fig. 6 presents stress distribution
The observation diagrams in Fig. 8 were prepared to at the height of the wall in the midspan cross section after
present coupled thermal-shrinkage stress distribution in 18.3 days under the assumption of uniform distribution

318 ACI Structural Journal/March-April 2014


Fig. 7—Stress distribution at height of wall in midspan cross section after 18.3 days under assumption of real (nonuniform)
distribution of: (a) temperature and (b) moisture content in wall; (c) thermal and (d) shrinkage stresses in the interior of the
wall; and (e) thermal and (f) shrinkage stress in interior and on surface of 20 m (65.6 ft) long wall.
of temperature in the wall, while in Fig. 7 and 8, the real, to the nonuniform distribution of temperature in the wall.
nonuniform distribution of temperature is taken into account. The importance of the self-induced stresses is discussed in
It accounts for the visible differences in the obtained stress a following section. It is interesting that the same differ-
distribution, especially in thermal stress near the joint. ences in the stress distribution can be obtained with the use
In this case, the self-induced stress arises in the wall due

ACI Structural Journal/March-April 2014 319


Fig. 8—Coupled thermal-shrinkage stress distribution at height of wall in midspan cross section after 18.3 days under assump-
tion of real (nonuniform) temperature and moisture content distribution in wall: (a) total stress in interior of wall; and (b)
stresses in interior and on surface of 20 m (65.6 ft) long wall.
in the wall. Total stresses exerted in the wall supported on
a stiff foundation due to the thermal-shrinkage effects are
a sum of self-induced and restraint stresses. To assess the
contribution of self-induced stresses in the analyzed walls,
the impact of restraint in a form of foundation was mini-
mized by reduction of the foundation’s stiffness to EF =
100 MPa (14.5 ksi) (flexible foundation was assumed).
Diagrams in Fig. 9 present development of self-induced
(Fig. 9(a)) and total (Fig. 9(b)) stresses in time for the loca-
tion in which the maximum value of stress was observed for
the 20 m (65.6 ft) long walls of 70 and 40 cm (2.3 and 1.3 ft)
thickness, assuming that both walls were detained in the
formwork. Stress development was presented for one length
of the wall for better visibility, as the values are similar.
The resulting self-induced stresses reach relatively low
values compared with the total stresses. Moreover, their
character is different and is closer to the behavior of typical
massive concrete structures. In the first phase, the interior
of the wall is subjected to compression, while surface layers
are tensioned; in the second phase, stress body inversion is
observed. This behavior is especially visible in thicker walls;
temperature and moisture content differences at the thick-
ness of the thinner walls are smaller, so the resulting stresses
are of lower value. It is worth noting that the generation of
self-induced stresses is the cause of total stress difference in
different zones of the wall.
Fig. 9—Thermal-shrinkage stress development in time: (a) Figure 10 presents a comparison of the diagrams of
self-induced stresses; and (b) total stresses in 20 m (65.6 ft) self-induced (Fig. 10(a)) and total stress (Fig. 10 (b)) distri-
long wall. bution in the midspan cross section of the wall in both the
heating (after 1.2 days) and cooling phase (after 18.3 days).
of analytical methods when the nonuniform distribution of The character of the observed stresses is similar in each
temperature at the height of the wall is taken.19 wall, so the diagrams are presented on the example of one
wall (L_20,d_0.7) only. Massive concrete-like behavior can
Self-induced versus restraint stresses be observed in unrestrained walls, while the signs of total
In the next stage, the computations were made to evaluate stresses are the same at the whole height of the wall.
the share of self-induced stresses in total stresses exerted

320 ACI Structural Journal/March-April 2014


Fig. 10—(a) Self-induced and (b) total stress distribution at height of 20 m (65.6 ft) long, 70 cm (2.3 ft) thick wall in midspan
cross section in heating (ph_I) and cooling (ph_II) phase.

Fig. 11—Stress distribution in midspan cross section of 20 m


(65.6 ft) long, 70 cm (2.3 ft) thick wall after 18.3 days from
which formwork was removed after 3 and 7 days.
Influence of time of formwork removal on stress
distribution
Finally, an analysis was performed that investigated the
influence of early formwork removal (3 and 7 days after
concrete casting). Diagrams in Fig. 11 show stress devel- Fig. 12—Stress distribution in midspan cross section of 20 m
opment in time, while Fig. 12 shows stress distribution after (65.6 ft) long, 70 cm (2.3 ft) thick wall after 18.3 days from
18.3 days in the midspan cross section for the walls from which formwork was removed after 3 and 7 days.
which the formwork was removed after 3 and 7 days. Because
of the similar character, diagrams were also presented for CONCLUSIONS
one exemplary wall (L_20,d_0.7). If the wall is kept in the The problem with high temperatures arising during the
formwork long enough for the concrete to cool completely, hardening of concrete has been known since the 1930s,
the heat concentration in the interior of the wall leads to when dams were first built in the United States. Much effort
higher stress and possible first crack development in the has been focused on the creation of efficient methods for
internal parts of the wall (Fig. 9(b) and 10(b)). When form- mitigation of the negative effects of concrete curing in
work is removed in early phases of concrete curing, greater massive structures; this problem is well known in concrete
stresses are observed on the surface of the wall as a result elements with considerable thickness. Nevertheless, forma-
of rapid cooling of the wall surface, which may lead to first tion of cracks is also observed in medium-thick concrete
crack formation in the near-surface areas (Fig. 11 and 12). It elements, such as RC walls cast against an old set founda-
should be noted that in both cases, cracks may extend to the tion. Such walls can also be sensitive to early-age cracking
entire wall thickness. It is important to note that formwork of thermal and shrinkage origins. Control of thermal and
removal accelerates moisture loss near the surface, which shrinkage cracking in early-age concrete is of great impor-
is why a significant increase of tensile stresses is observed tance to ensure a desired service life and function of struc-
in the vicinity of the top surface if it is no longer protected. tures. It is a complicated problem due to the complex nature

ACI Structural Journal/March-April 2014 321


of interacting phenomena and a large number of contributing ACKNOWLEDGMENTS
factors. Important factors include the dimensions of struc- This paper was done as a part of a research Project N N506 043440 enti-
tled, “Numerical Prediction of Cracking Risk and Methods of Its Reduction
tural elements and the length-height ratio, which directly in Massive Concrete Structures,” funded by the Polish National Science
affect the level of a wall restraint in a foundation. Centre. The co-author of the paper, Agnieszka Knoppik-Wróbel, is a
The contribution of self-induced and restrained stresses scholar under the Project SWIFT, co-financed by the European Union under
the European Social Fund.
to total stresses induced in the wall with different dimen-
sions was investigated. The results obtained with the use of
REFERENCES
the original numerical model were discussed. Very limited 1. ACI Committee 207, “Report on Thermal and Volume Change Effects
analysis of the stress distribution in the walls could be found on Cracking of Mass Concrete (ACI 207.2R-07),” American Concrete Insti-
in the literature concerning early-age concrete. This paper tute, Farmington Hills, MI, 2007, 28 pp.
2. Branco, F. A.; Mendes, P. A.; and Mirambell, E., “Heat of Hydra-
is an attempt to fill the vacancy in this field, in particular by tion Effects in Concrete Structures,” ACI Materials Journal, V. 89, No. 2,
providing information about development of stresses and the Mar.-Apr. 1992, pp. 139-145.
importance of self-induced stresses in externally restrained 3. De Schutter, G., “Fundamental Study of Early Age Concrete Behaviour
as a Basis for Durable Concrete Structures,” Materials and Structures,
structures. Knowledge about the distribution of stresses is V. 35, Jan.-Feb. 2002, pp. 15-21.
necessary in many practical cases, for example, in the eval- 4. Flaga, K., and Furtak, K., “Problem of Thermal and Shrinkage
uation of crack risk of early-age concrete structures. The Cracking in Tanks Vertical Walls and Retaining Walls Near Their Contact
with Solid Foundation Slabs,” Architecture–Civil Engineering–Environ-
results of the analysis can be summarized as follows: ment, V. 2, No. 2, June 2009, pp. 23-30.
1. Thermal stresses play a predominant role in the total 5. Zych, M., “Analiza pracy scian zbiorników zelbetowych we wczesnym
thermal-shrinkage stress development; okresie dojrzewania betonu w aspekcie ich wodoszczelnosci (Analysis of
Work of RC Tank Walls in Early Ages of Concrete Curing in the View of
2. The total thermal-shrinkage stresses arising in an RC Their Water Tightness),” PhD thesis, Faculty of Civil Engineering, Cracow
wall result mainly from restraint stresses generated by Technical University, 2011. (in Polish)
limited possibility of the wall deformation; the share of 6. Benboudjema, F., and Torrenti, J. M., “Early-Age Behaviour of
Concrete Nuclear Containments,” Nuclear Engineering and Design, V.
self-induced stresses increases with the increasing thickness 238, No. 10, Oct. 2008, pp. 2495-2506.
of the wall, but even in relatively thin elements, it has an 7. Estrada, C. F.; Godoy, L. A.; and Prato, T., “Thermo-Mechanical
influence on total stress distribution in the wall; and Behaviour of a Thin Concrete Shell during Its Early Age,” Thin-Walled
Structures, V. 44, No. 5, May 2006, pp. 483-495.
3. Three-dimensional numerical analysis results explain 8. De Borst, R., and Van den Boogaard, A. H., “Finite Element Modeling
the following phenomena observed in externally restrained of Deformation and Cracking in Early-Age Concrete,” Journal of Engi-
elements: neering Mechanics, ASCE, V. 120, No. 12, Dec. 1994, pp. 2519-2534.
9. Nilsson, M., “Restraint Factors and Partial Coefficients for Crack Risk
• The greatest thermal-shrinkage stress does not occur Analyses of Early Age Concrete Structures,” PhD thesis, Department of
at the interface between the wall and the restraint but Civil and Mining Engineering, Luleå University of Technology, Sweden,
at some level above the restraint joint. That fact results 2003.
10. Larson, M., “Thermal Crack Estimation in Early Age Concrete.
from nonuniform distribution of temperature and mois- Models and Methods for Practical Application,” PhD thesis, Department of
ture within the element which concentrate in its central Civil and Mining Engineering, Luleå University of Technology, Sweden,
parts. In this case, the self-induced stress is also consid- 2003.
11. ACI Committee 207, “Guide to Mass Concrete (ACI 207.1R-05),”
ered. The first crack can consistently be observed at American Concrete Institute, Farmington Hills, MI, 2005, 30 pp.
some level above the restraint joint; and 12. RILEM TC 119-TCE, “Recommendations of TC 119-TCE: Avoid-
• When the wall is detained in formwork until it cools ance of Thermal Cracking in Concrete at Early Ages,” Materials and Struc-
tures, V. 30, Oct. 1997, pp. 451-464.
down, stresses develop towards the interior of the wall; 13. RILEM Report 25, “Early Age Cracking in Cementitous Systems,”
thus, internal cracking may initially develop. When Final Report of RILEM Technical Committee TC 181-EAS, 2002.
formwork is removed in early phases of concrete curing, 14. Mihashi, H., and Leite, J. P., “State-of-the-Art Report on Control
Cracking in Early Age Concrete,” Journal of Advanced Concrete Tech-
an increased cooling rate leads to greater stresses in nology, V. 2, No. 2, June 2004, pp. 141-154.
the surface zones, and first cracks can develop on the 15. Majewski, S., “MWW3—Elasto-Plastic Model for Concrete,”
surface of the wall. Archives of Civil Engineering, V. 50, No. 1, 2004, pp. 11-43.
16. Klemczak, B., “Adapting of the Willam-Warnke Failure Criteria for
Young Concrete,” Archives of Civil Engineering, V. 53, No. 2, June 2007,
AUTHOR BIOS pp. 323-339.
Barbara Klemczak is an Associate Professor in the Department of Civil 17. Klemczak, B., “Prediction of Coupled Heat and Moisture Transfer in
Engineering at the Silesian University of Technology, Gliwice, Poland. Early-Age Massive Concrete Structures,” Numerical Heat Transfer. Part A:
She received her PhD and DSc from the Silesian University of Technology Applications, V. 60, No. 3, Sept. 2011, pp. 212-233.
in the field of numerical modeling of early-age massive concrete. Her 18. Comité Euro-International du Béton, “CEB-FIP Model Code 1990,”
research interests include nonlinear analysis of reinforced concrete struc- Thomas Telford, London, UK, 1991, 437 pp.
tures, particularly numerical modeling of thermal and shrinkage effects in 19. Klemczak, B., and Knoppik-Wróbel, A., “Comparison of Analytical
concrete structures at early ages. Methods for Estimation of Early-Age Thermal-Shrinkage Stresses in RC
Walls,” Archives of Civil Engineering, V. 59, No. 1, 2013, pp. 97-117.
Agnieszka Knoppik-Wróbel is a PhD Student in the Department of Civil
Engineering at the Silesian University of Technology. Her research interests
include cracking risk in early-age externally restrained concrete structures.

322 ACI Structural Journal/March-April 2014


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 111-S28

Effects of Casting Position and Bar Shape on Bond of


Plain Bars
by Montserrat Sekulovic MacLean and Lisa R. Feldman
Twenty-five lap splice specimens were reinforced with plain round This paper presents the design and results of an experi-
or square longitudinal bars in the top or bottom position to eval- mental program to evaluate the effects of casting position
uate the effects of casting position and bar shape on bond. All spec- and bar shape on plain steel bars longitudinally cast in lap
imens failed in bond, and the bond of square bars may be evaluated splice specimens.
by calculating their equivalent round diameter. Top cast factors of
0.3 and 0.6 for round and square bars, respectively, reasonably
RESEARCH SIGNIFICANCE
capture the reductions in bond resistance. Maximum load predic-
tions based on the CEB-FIP draft Model Code 2010 provisions for Both round and square plain steel reinforcing bars are
bond are overly conservative for all combinations of bar shape regularly encountered in historical structures, and criteria for
and casting position, whereas CEB-FIP Model Code 1990 provi- assessing their bond strength are necessary. The results of
sions for bond reasonably and conservatively capture the behavior an experimental investigation of lap splice specimens rein-
of specimens with bottom-cast round bars, but do not appear to forced with plain bars are presented to evaluate the effects
capture the behavior of specimens with bottom- or top-cast square of bar shape and casting position on bond behavior. Such
bars or top-cast round bars. knowledge will aid in the development of bond provisions
for the evaluation of concrete structures reinforced with
Keywords: bar shape; bond; casting position; lap splice; plain reinforce-
ment; stress.
plain steel bars.

INTRODUCTION EXPERIMENTAL INVESTIGATION


Recently published works highlight case studies of The description of the specimens, materials, and test setup
historic reinforced concrete structures with plain reinforce- are similar to those described by Hassan and Feldman,7 but
ment,1 and reviews of structural inventories include many are briefly described herein for comprehensiveness. Figure
structures with reinforcing bar details and types that do not 1 shows the cross sections, elevation, and plan view for
meet current requirements.2 Many concrete structures with the 25 specimens in this study. Ten of the specimens, as
plain reinforcement have reached an age where they require identified in Table 1, were originally reported by Hassan
remediation,3,4 and so it is crucial for forensic engineers to and Feldman.7 All specimens had identical cross-sectional
have an understanding of their behavior and capacity. Of dimensions and span lengths. Figure 1(a) and (b) show the
particular note, ACI Committee 562, organized in 2004, has cross section of specimens with the round or square longi-
a goal of developing a code and commentary for the evalu- tudinal reinforcement cast in the bottom and top positions,
ation, repair, and rehabilitation of existing concrete struc- respectively. These cross sections show that the cover was
tures, and will likely need to include provisions for the bond held constant at 50 mm (2 in.) regardless of the size of the
evaluation of plain reinforcement. longitudinal reinforcing bars used in the various specimens.
Plain reinforcement does not possess lugs or other surface Earlier works2,8 suggested that the bond strength of plain
deformations, and cannot transfer bond forces by mechan- reinforcement is independent of concrete cover because
ical interlock. Instead, bond is transferred through adhe- these bars lack mechanical interlock with the surrounding
sion between the concrete and the reinforcement before slip concrete; thus, the likelihood of a splitting failure is reduced.
occurs, and by wedging of small particles that break free Specimens cast with top reinforcement were inverted before
from the concrete upon slip.5 Moreover, both plain square testing such that Fig. 1(c) shows the elevation of all spec-
and round bars have been used to reinforce concrete struc- imens as tested, including the span length, loading, and
tures,6 though only round bars were included in Abrams’ reinforcing steel arrangement. The shear span-depth ratio,
historic study.5 The extent of void formations beneath a/d, was approximately equal to 3.94 for all specimens.
top-cast round and square bars due to the upward migra- Figure 1(d) shows a plan view of the specimens and illus-
tion of water and mortar that occurs during concrete place- trates the arrangement of the spliced longitudinal bars.
ment operations, and differences in concrete consolidation All specimens were designed to fail in bond, and had lap
around the two bar shapes, might cause different relation- splice lengths Ls ranging from 12.8 to 32.1 times the longi-
ships between the required lap splice length and the depth tudinal bar diameter for round bars or the side face dimen-
of concrete cast under the bar. Plain bars may also be more ACI Structural Journal, Vol. 111, No. 2, March-April 2014.
affected by casting position than deformed bars because the MS No. S-2012-097.R1, doi:10.14359.51686524, was received August 3, 2012, and
reviewed under Institute publication policies. Copyright © 2014, American Concrete
adhesion component of bond is a more dominant factor in Institute. All rights reserved, including the making of copies unless permission is
obtained from the copyright proprietors. Pertinent discussion including author’s
the transfer of bond force for plain bars. closure, if any, will be published ten months from this journal’s date if the discussion
is received within four months of the paper’s print publication.

ACI Structural Journal/March-April 2014 323


Table 1—Actual and predicted failure loads
Predicted normalized maximum load Pmax /√fc′, kN/√MPa (lb/√psi)
Maximum
Splice length Concrete Bar surface normalized Neglecting
as a function of compressive roughness Ry, load, Pmax /√fc′, strain Including strain CEB-FIP Draft CEB-FIP
Specimen bar size strength fc′, μm kN/√MPa hardening of hardening of Model Code Model Code
identification* (Ls/db) MPa (psi) (× 10–3 in.) (lb/√psi) reinforcement reinforcement 1990 2010
19l-305↓† 16.1 17.4 (2520)‡ 9.54 (0.376) 8.50 (159) 18.0 (337) 29.1 (544) 4.98 (93.0) 0.96 (17.9)
† ‡
19l-410↓ 21.6 17.4 (2520) 9.67 (0.381) 9.14 (171) 18.0 (337) 29.1 (544) 7.72 (144) 2.33 (43.5)
† ‡
19l-510↓ 26.8 18.7 (2710) 9.86 (0.388) 9.58 (179) 17.5 (327) 28.4 (530) 10.5 (196) 3.71 (69.2)
† ‡
19l-610↓ 32.1 21.0 (3040) 9.44 (0.372) 17.8 (332) 16.7 (311) 27.1 (507) 13.9 (259) 5.36 (100)
† ‡
25l-410↓ 16.4 23.7 (3440) 8.88 (0.350) 16.2 (302) 28.1 (524) 45.0 (840) 12.5 (233) 4.12 (76.9)
† ‡
25l-510↓ 20.4 24.0 (3480) 8.43 (0.332) 18.4 (343) 27.7 (518) 44.5 (831) 16.1 (300) 5.74 (107)
† ‡
25l-610↓ 24.4 22.8 (3310) 8.71 (0.343) 20.6 (384) 28.5 (532) 45.6 (851) 19.8 (370) 7.52 (140)
§
25l-410↑ 16.4 27.1 (3930) 9.19 (0.362) 6.55 (122) 28.1 (524) 42.2 (788) 8.39 (157) 0.91 (17.0)
25l-510↑ 20.4 28.0 (4060)§ 9.09 (0.358) 4.69 (87.5) 27.7 (517) 41.7 (778) 11.1 (207) 1.76 (32.8)
25l-610↑ 24.4 35.8 (5190)§ 9.21 (0.362) 7.07 (132) 25.1 (468) 38.0 (710) 14.8 (276) 2.91 (54.3)
32l-410↓† 12.8 19.8 (2870)‡ 9.92 (0.390) 15.6 (291) 44.5 (827) 63.2 (1180) 14.1 (263) 4.43 (82.7)
32l-610↓† 19.1 19.8 (2870)‡ 9.72 (0.383) 25.1 (468) 44.3 (827) 63.2 (1180) 22.1 (412) 7.94 (148)
32l-810↓† 25.3 15.8 (2290)‡ 10.1 (0.398) 31.8 (594) 46.9 (876) 64.0 (1190) 28.0 (523) 10.9 (203)
§
25n-410↓ 16.4 25.5 (3700) 8.79 (0.346) 16.1 (300) 38.3 (714) 55.3 (1030) 14.8 (276) 4.85 (90.5)
§
25n-510↓ 20.4 25.0 (3620) 8.83 (0.348) 20.0 (373) 38.3 (714) 55.2 (1031) 18.6 (347) 6.50 (121)
§
25n-610↓ 24.4 28.1 (4080) 8.86 (0.349) 26.8 (500) 36.6 (684) 53.4 (996) 22.9 (427) 8.24 (154)
§
25n-410↑ 16.4 33.0 (4790) 9.06 (0.357) 8.97 (167) 31.3 (585) 50.2 (937) 10.5 (196) 1.47 (27.4)
§
25n-510↑ 20.4 33.5 (4860) 9.17 (0.361) 11.2 (209) 31.1 (581) 49.9 (932) 13.6 (254) 2.37 (44.2)
§
25n-610↑ 24.4 33.0 (4790) 9.16 (0.361) 12.2 (228) 31.3 (585) 50.2 (937) 16.6 (310) 3.26 (60.8)
§
32n-410↓ 12.8 25.5 (3700) 9.36 (0.368) 17.4 (325) 50.6 (944) 73.7 (1380) 17.3 (323) 5.69 (106)
§
32n-610↓ 19.1 25.5 (3700) 9.24 (0.364) 20.1 (375) 50.8 (948) 74.0 (1380) 26.6 (496) 9.63 (180)
§
32n-810↓ 25.3 26.9 (3900) 9.17 (0.361) 28.3 (528) 49.5 (924) 72.9 (1360) 35.6 (664) 13.4 (250)
§
32n-410↑ 12.8 27.5 (4000) 9.29 (0.366) 12.6 (235) 49.4 (922) 73.0 (1360) 11.6 (216) 1.66 (31.0)
§
32n-610↑ 19.1 26.2 (3800) 9.52 (0.375) 14.3 (267) 50.4 (941) 73.8 (1380) 17.9 (334) 3.54 (66.1)
32n-810↑ 25.3 26.2 (3800)§ 9.34 (0.368) 16.2 (302) 50.4 (941) 73.8 (1380) 24.5 (457) 5.50 (103)
*
First number in specimen identification represents nominal diameter for round bars or side face dimension for square bars. Solid circle (l) or square (n) identifies shape of longi-
tudinal reinforcement. Number following hyphen denotes lap splice length, in millimeters, with an up arrow (↑) showing that longitudinal bars were cast in top position, and down
arrow (↓) showing that bars were cast in bottom position.

Originally reported by Hassan and Feldman (2012).

Specimens cast with Concrete Mixture Proportion 1.
§
Specimens cast with Concrete Mixture Proportion 2.
sion for square bars. Specimen dimensions and lap splice proportion used to cast each specimen. Mixture Proportion
lengths were selected to match those reported by Idun and 1 consisted of a crushed limestone and granite coarse aggre-
Darwin9 for tests of specimens with deformed bars. A direct gate blend and a silica sand fine aggregate. The mixture
comparison of the specimens with plain and deformed bars proportion per cubic meter (cubic yard) of concrete was:
is reported elsewhere.7 250 kg (421 lb) cement, 1100 kg (1854 lb) sand, 1100 kg
Figure 2 shows that a vertical load was applied using a (1854 lb) crushed coarse aggregate, and 140 L (28.3 gal.)
spreader beam with a self-weight that exerted a load P of water. Mixture Proportion 2 consisted of a carbonate,
1.77 kN on the specimens to establish the four-point loading gneiss, and granite coarse aggregate blend and a washed
arrangement. Loading was applied at a rate of 0.0015 mm/s silica sand fine aggregate. The mixture proportion per cubic
(0.0006 in./s) to failure. meter (cubic yard) of concrete was: 270 kg (455 lb) cement,
993 kg (1674 lb) sand, 1039 kg (1751 lb) crushed coarse
Concrete aggregate, and 145 L (29.3 gal.) water. The maximum size
The concrete had a target compressive strength of 20 MPa of the coarse aggregate used in both mixture proportions
(2900 psi). General purpose (Type GU) portland cement was 20 mm (0.8 in.), and all aggregates conformed to CAN/
was used without admixtures. Table 1 indicates the mixture CSA A23.1-09.10 It should be noted that the effect of mixture

324 ACI Structural Journal/March-April 2014


Fig. 1—Splice specimen geometry: (a) cross section for specimens with bottom-cast longitudinal reinforcement; (b) cross
section for specimens with top-cast reinforcement; (c) elevation; and (d) plan view. (Note: Dimensions are given in mm [in.].)

Fig. 2—Test setup.

ACI Structural Journal/March-April 2014 325


Table 2—Longitudinal reinforcing steel properties
Static yield strength fys, Dynamic yield strength fyd, Ultimate strength fu, Modulus of elasticity Es,
Bar identification MPa (ksi) MPa (ksi) MPa (ksi) GPa (ksi)
19l↓ 326 (47.3) 355 (51.5) 520 (75.4) 203 (29,400)
25l↓ 322 (46.7) 346 (50.2) 534 (77.4) 196 (28,400)
25l↑ 340 (49.3) 364 (52.8) 522 (75.7) 243 (35,200)
32l↓ 318 (46.1) 348 (50.5) 504 (73.1) 204 (29,600)
25n↓ 357 (51.8) 381 (55.2) 544 (78.9) 192 (27,800)
25n↑ 325 (47.1) 349 (50.6) 542 (78.6) 207 (30,000)
32n↓ and 32n↑ 320 (46.4) 343 (49.7) 527 (76.4) 196 (28,400)

proportion on bond strength was not within the scope of the would be governed by bond between the longitudinal rein-
current investigation. Rather, the change in mixture propor- forcement and the surrounding concrete.
tions resulted from a change in ready-mix suppliers. Bleed
water measurements were not conducted as part of this EXPERIMENTAL RESULTS
investigation. Table 1 shows the observed maximum loads attained
Table 1 shows the concrete compressive strength of the by the specimens and those predicted assuming yielding
specimens at the time of testing as established from the of the reinforcement, both neglecting and including strain
results of companion concrete cylinders stored under the hardening, and predicted loads based on average bond
same conditions and tested on the same day as the corre- stress provisions for plain reinforcement included in the
sponding splice specimen. Specimens were moist cured CEB-FIP Model Code 199014 and the CEB-FIP draft Model
using wet burlap and plastic sheets for 7 days following Code 2010.15 The predicted loads have been reduced by the
casting, and were then stored in the laboratory until testing. weight of the spreader beam and the specimen self-weight to
allow for a direct comparison with the maximum loads that
Reinforcement were recorded during testing. All reported loads have been
All principal longitudinal reinforcement was hot-rolled normalized by the square root of the concrete compressive
CSA G40.21 300W steel. Figure 1(c) shows that the bars strength given that a previous work8 showed that it is valid
had 180-degree hooks at the ends adjacent to the beam for plain reinforcement and is consistent with familiar equa-
supports to ensure that the bond failure occurred within the tions for deformed bars.
lap splice length. The material properties were established The specimens are identified by mark numbers that include
from coupons obtained from surplus bar lengths and tested two numbers and associated symbols separated by a hyphen.
in accordance with ASTM A370.11 Table 2 shows the static The first number represents the nominal diameter for round
yield strength fys calculated in accordance with Rao et al.,12 bars or the nominal side face dimensions for square bars db,
dynamic yield strengths fyd, ultimate strength fu, and modulus in millimeters, used to longitudinally reinforce the speci-
of elasticity Es for all longitudinal bar sizes used. mens. A solid circle (l) or square (n) following this number
The longitudinal reinforcing bars were sandblasted using identifies the shape of the longitudinal reinforcement. The
220-grit aluminum oxide, a nozzle distance of 125 mm number following the hyphen denotes the lap splice length
(5 in.), and a blast pressure of 698 kPa (100 psi) to increase Ls in millimeters, with an up arrow (↑) showing that the
the surface roughness and make them more representative of longitudinal bars were cast in the top position (Fig. 1(b)),
historical bars.8 The surface roughness of each bar was char- or a down arrow (↓) showing that the bars were cast in the
acterized by the maximum height of profile Ry, established bottom position (Fig. 1(a)).
as the distance between the highest peak and the deepest Table 1 shows that the maximum normalized load reported
valley on the bar surface.13 Table 1 shows the average Ry for Specimen 25l-510↑ was only 72% that of Spec-
values for the longitudinal reinforcing bars in each specimen imen  25l-410↑, a specimen cast from the same batch of
based on a total of 30 roughness measurements on each concrete. This result was considered suspect because speci-
bar using a surface roughness tester and a single 0.25 mm mens with longer splice lengths should be able to resist higher
(0.01 in.) stroke. loads when all other variables are held constant. Removal
The shear reinforcement consisted of 12.7 mm (0.5 in.) of the concrete surrounding the longitudinal reinforcement
diameter hot-rolled CSA G40.21 300W plain steel bars was completed for Specimen 25l-510↑ following testing;
spaced at 200 mm (8 in.) on center within the shear spans, however, no voids were identified that would have impaired
and 250 mm (10 in.) on center within the constant moment the bond between the reinforcement and the surrounding
region outside of the lap splices (Fig. 1(d)). Two additional concrete. This specimen therefore could not be identified as
stirrups were placed in the splice region one-quarter of the a physical outlier, and is included in the regression analysis
splice length, but not exceeding 150 mm (6 in.), from the as will be presented in a subsequent section.
ends of the splice to prevent prying action of the longitudinal Table 1 shows that all but Specimen 19l-610↓ failed at
reinforcement. The specimens had considerably more shear loads well below those predicted using the flexural resis-
reinforcement than strictly necessary to ensure that failure tance procedures in ACI 318-1116 with resistance factors set

326 ACI Structural Journal/March-April 2014


equal to unity, and thus suggest that these specimens failed The bar size factor η3 in Eq. (1) and (2) was therefore
in bond. The load-deflection behavior and crack patterns for calculated assuming db = db,EQ for the case of specimens
all specimens is consistent with that described by Hassan longitudinally reinforced with square bars.
and Feldman,7 including that of Specimen 19l-610↓, which, The predicted maximum normalized loads for all speci-
based upon its load-deflection behavior, failed in bond even mens tested were then calculated assuming a linear strain
though it attained a maximum load of 106% that predicted distribution along the height of the specimen and the stress
assuming yielding of the reinforcement. versus strain distribution for the concrete as obtained from
companion specimens tested in conjunction with each
Comparison of test results with predictions based splice specimen. The neutral axis location was established
on European code provisions from cracked transformed section properties because all
Both the CEB-FIP Model Code 199014 and the draft Model specimens failed in bond, and, with the exception of Spec-
Code 201015 provide equations for the average bond stress imen 19l-610↓, at loads well below those predicted based
for plain reinforcement, and thus allow for a prediction of on yielding of the longitudinal reinforcement.
the maximum normalized load resisted by the specimens in Figure 3(a) shows the ratio of the test-to-predicted
the study, as reported in Table 1. The design average bond maximum loads based on the CEB-FIP Model Code 199014
stress uave for reinforcing bars is specified in the CEB-FIP provisions, Pmax /(Pmax)CEB,1990, for all specimens. Values of
Model Code 199014 as Pmax /(Pmax)CEB,1990 > 1.0 suggest that the predicted values
are conservative, whereas values of Pmax /(Pmax)CEB,1990 < 1
uave = η1η2 η3 fctd (1) suggest that the same values are unconservative. The average

Pmax /(Pmax)CEB,1990 for all specimens with bottom-cast round
where η1 is a factor that accounts for the reinforcing type, bars is equal to 1.2, and so suggests that the CEB-FIP Model
Code 199014 provisions reasonably and conservatively
and is equal to 1.0 for plain bars; η2 considers the casting
capture the bond behavior of these bars. The same code
position of the reinforcement, and is equal to 1.0 and 0.7
provisions, however, do not appear to capture the behavior of
for bottom and top cast reinforcement, respectively; η3
specimens with bottom- (average Pmax /(Pmax)CEB,1990 = 0.98)
considers the bar size, and is equal to 1.0 for bars with a
or top-cast square bars (average Pmax /(Pmax)CEB,1990 = 0.83),
diameter db of 32 mm (1.25 in.) or less and (132 – db)/100 for
or specimens reinforced with top-cast round bars (average
db > 32 mm (1.25 in.); and fctd is the design tensile strength
Pmax /(Pmax)CEB,1990 = 0.56).
of the concrete in MPa.
Figure 3(b) shows that the ratio of the test-to-predicted
Bond provisions have changed markedly in the CEB-FIP
loads based on the CEB-FIP draft Model Code 201015
draft Model Code 2010.15 The basic average bond strength
provisions, Pmax /(Pmax)CEB,2010, for all specimens. All values
uave, which can be used to assess plain reinforcement, is
of Pmax /(Pmax)CEB,2010 exceed 2.0, which suggests that the
CEB-FIP draft Model Code 201015 provisions are overly
uave = η1η2 η3 η4 ( fck 20 )
0.5
γc (2)
conservative. Figure 3(b) does suggest that these provisions
are also more conservative when estimating the capacity of
where η1 is equal to 0.9 for plain bars; η2 is equal to 1.0 specimens cast with the longitudinal reinforcement in the
and 0.5 for plain bars cast in the bottom and top positions, top position, and that the provisions tend to be more conser-
respectively; η3 is equal to 1.0 for db ≤ 20 mm (0.79 in.) vative for specimens with shorter lap splice lengths for both
and (20/db)0.3 for larger bar sizes; η4 accounts for the charac- bar shapes and casting positions.
teristic yield stress of the reinforcement and is equal to 1.2
for reinforcement with a yield stress equal to, and presum- Top casting effects
ably less than, 400 MPa (58.0 ksi); fck is the characteristic Figure 4 shows the ratio of the maximum normalized
value of the cylinder compressive strength of the concrete, in loads for each pair of specimens for which the same size,
MPa; and γc is a partial factor for the concrete compressive shape, and lap splice length were provided for longitudinal
strength set equal to unity for the case of using Eq. (2) as a reinforcement, but with this reinforcement cast in the top
predictive, rather than design, equation. position for one specimen, and cast in the bottom position
Neither edition of the CEB-FIP Model Code14,15 specif- for the other specimen. Figure 4 shows that all specimens
ically provides for square reinforcing bars. An analysis of with top cast reinforcement failed at loads well below those
archival test results of pullout specimens reinforced with for bottom cast reinforcement, with a resulting average
historical round and square bars whose deformation patterns ratio of the normalized maximum loads equal to 0.51. The
did not conform to ASTM A305-4917 was performed by results of this limited investigation suggest that square rein-
Howell and Higgins18 and showed that the simplified ACI forcement, with an average ratio of 0.60, is less sensitive to
development length equations19 provided a lower bound for casting position than round bars, with a resulting average
both bar shapes, where square bars were evaluated by calcu- ratio of 0.33. Furthermore, the larger square reinforcing bar
lating their equivalent round diameter db,EQ size, db = 32 mm (1.25 in.), used to longitudinally reinforce
the specimens, appears to be less sensitive to casting posi-
4 db 2 tion than the smaller square bars with db = 25 mm (1 in.).
db, EQ = (3)
π Current American16 and Canadian20 code provisions for
reinforced concrete require a 30% increase in development

ACI Structural Journal/March-April 2014 327


Fig. 4—Ratio of normalized maximum loads for pairs of
specimens with longitudinal reinforcement in top versus
bottom cast position.
below the bottom face of the bar only and will affect a much
smaller portion of the overall perimeter.

Effect of bar shape as assessed using predictive


equations for maximum normalized load
A regression analysis of the 25 specimens yields
the following empirical equation for the normalized
maximum load
Fig. 3—Test-to-predicted ratio of maximum normalized
loads: (a) predicted loads based on CEB-FIP 1990 code
Pmax Ls
provisions; and (b) predicted loads based on CEB-FIP 2010 = 9.38 × 10 −5 Ls db Ry + (0.24 + 0.15kb − 0.50kc ) (4)
draft code provisions. fc′ db

length for deformed bars cast in the top position (that is, where Ls is the lap splice length, in mm; db is the longitu-
when more than 300 mm [12 in.] of fresh concrete is placed dinal bar size, in mm, reported as the measured bar diam-
below the reinforcement), whereas the CEB-FIP Model eter for round bars or the measured side face dimension
Code 199014 provides a multiplier of 0.7 to the average bond for square bars; Ry is the surface roughness of the longitu-
stress in such cases. The modifiers in all three of these codes dinal reinforcement, in μm; kb is an indicator variable for
appear to be unconservative for plain bars in light of the the shape of the longitudinal bars, and is equal to zero for
results discussed in the previous paragraph. Furthermore, a round bars and 1 for square bars; and kc is an indicator vari-
previous investigation conducted by Chana21 also concludes able for the casting position of the longitudinal reinforce-
that the bond of plain bars is more affected by casting posi- ment, and is equal to 0 if the bars are cast in the bottom
tion than that of deformed bars. This conclusion appears position (Fig. 1(a)), and 1 if the bars were cast in the top
justified upon consideration of the mechanics of bond. The position (Fig. 1(b)). The root mean square error for Eq. (4) is
adhesion between the concrete and the reinforcement is a 2.91 kN/√MPa (54.3 lb/√psi).
more dominant factor in the transfer of bond forces for plain Using an equivalent round diameter db,EQ, as described
reinforcement, because they cannot transfer these forces by by Eq. (3), results in the following predictive equation that
mechanical interlock.22 allows for the elimination of the indicator variable kb
In contrast, the CEB-FIB draft Model Code 201015
provides a multiplier of 0.5 specifically for plain bars cast Pmax
in the top position, and appears to be more reasonable when = 3.36 × 10 −4 Ls Ry db, EQ 0.5 (2.12 − kc ) (5)
fc′
compared with the results in the current investigation. The
test results presented herein, however, also show that square
bars are less sensitive to casting position than round bars where db,EQ is the diameter for round bars and the equiva-
due to differences in the shape of voids that form beneath lent round diameter of square bars. The resulting root-mean-
these two bar shapes. For round bars, the concrete tends to square error of 3.01 kN/√MPa (56.2 lb/√psi) for Eq. (5)
settle such that a void forms under the bottom half of the is similar to that reported for Eq. (4), and suggests that
perimeter23; whereas for square bars, assuming construction using the equivalent round diameter for plain square bars
allows them to be placed perfectly square within the rein- is reasonable.
forcing cage as shown in Fig. 1(a) and (b), the void will form Results of a previous investigation8 have shown that the
average surface roughness, Ry = 9.26 μm, for the 25 speci-
mens included in the regression analysis is a lower bound for

328 ACI Structural Journal/March-April 2014


Fig. 5—Comparison of recorded normalized maximum loads to those predicted empirically using Eq. (6) for: (a) specimens
cast with round longitudinal bars; and (b) specimens cast with square longitudinal bars.
that of historical bars. Furthermore, the analysis presented in
the previous section suggests that multipliers of 0.3 and 0.6
for round and square longitudinal bars, respectively, reason-
ably capture the reduction in bond resistance provided by
bars in the top cast position. The following simplified predic-
tive normalized maximum load equation therefore results

Pmax
= 6.63 × 10 −3 Ls ψ t db, EQ 0.5 (6)
fc′

where ψt is a factor used to modify the lap splice length


based on the location of the longitudinal reinforcement, and
is equal to 1.0 for both round and square bars cast in the
bottom position, 0.3 for round bars cast in the top position,
and 0.6 for square bars cast in the top position. The resulting
root-mean-square error for Eq. (6) is 2.44 kN/√MPa Fig. 6—Predicted versus recorded maximum normalized
(45.5 lb/√psi). loads.
Figure 5 shows the fit of Eq. (6) with the experimental test
cast with square longitudinal bars cast in the top position
data, with Fig. 5(a) showing all data for specimens longitu-
attained a mean predicted-recorded maximum normalized
dinally reinforced with round bars, and Fig. 5(b) showing
load ratio of 1.02, and a standard deviation of 0.151. Results
all data for specimens longitudinally reinforced with square
for specimens longitudinally reinforced with round bars in
bars. It should be noted that a linear and proportional rela-
the bottom position had a mean predicted-recorded load
tionship, with Pmax /√fc′ = 0 for Ls = 0, is the best fit. This
ratio of 0.983, and a standard deviation of 0.159. In contrast,
finding differs from the linear, but not proportional, relation-
specimens longitudinally reinforced with round bars cast in
ship reported for deformed bars16 and prestressing strands.24
the top position attained a mean value and standard devia-
An alternate method of evaluating the goodness of fit of
tion for the predicted-recorded applied load ratios of 1.22
Eq. (6) to the test data is presented in Fig. 6, which shows
and 0.340, respectively. The higher standard deviation
the predicted normalized load calculated in accordance
obtained for specimens cast with round longitudinal bars in
with Eq. (6) versus the recorded maximum normalized
the top position is likely due to the limited number of such
load for all 25 specimens in the test database. Also shown
specimens in the test database and the sensitivity of round
is the proportional line, which represents the theoretical
bars to casting position, as outlined in the previous section.
case that the predicted maximum normalized load exactly
equals that recorded during specimen testing. Specimen
SUMMARY AND CONCLUSIONS
markers falling above this line represent cases for which the
Twenty-five lap splice specimens with a shear span-depth
maximum normalized load, predicted in accordance with
ratio approximately equal to 3.94 were reinforced with plain
Eq. (6), exceeds the maximum normalized load recorded
round or square longitudinal steel bars in the top or bottom
during specimen testing. Similarly, markers falling below
position to evaluate the effects of casting position and bar
the proportional line represent cases for which Eq. (6) under-
shape on bond. Specimens were 305 mm (12 in.) wide by
estimates the maximum normalized load recorded during
410 mm (16 in.) tall, with a span length of 4570 mm (15 ft),
specimen testing.
and were subjected to four-point loading. Lap splice lengths
A review of the data presented in Fig. 6 shows that the
ranged from 12.8 to 32.1 times the longitudinal bar diam-
mean and standard deviation obtained for specimens longi-
eter or side face dimension for the case of round and square
tudinally reinforced with square bars cast in the bottom posi-
longitudinal bars, respectively. The following significant
tion were 1.04 and 0.154, respectively. Similarly, specimens
observations and conclusions were noted:

ACI Structural Journal/March-April 2014 329


• All specimens failed due to bond loss between the longi- η2 = factor to account for bond conditions
η3 = factor to account for bar size
tudinal reinforcement and the surrounding concrete; η4 = factor to account for characteristic yield strength of reinforcement
• Square longitudinal bars may be evaluated by calcu- ψt = modification factor for bar shape and casting position
lating their equivalent round diameter, based on equal l, n = symbols identifying bar shape
↑, ↓ = symbols identifying casting position of reinforcement
cross-sectional areas of the actual square reinforcing bar
and the equivalent round bar;
• Predictions of the maximum applied load based on REFERENCES
1. Feldman, L. R.; MacFarlane, D. C.; Kroman, J. A.; and Bartlett, F.
CEB-FIP Model Code 1990 provisions for bond reason- M., “Construction Staging of the Centre Street Bridge Rehabilitation to
ably and conservatively capture the behavior of specimens Accommodate Emergency Vehicle Traffic,” 31st Annual Conference of the
Canadian Society for Civil Engineering, 2003, 10 pp. (CD-ROM)
with bottom cast round bars, but do not appear to capture 2. Baldwin, M. I., and Clark, L. A., “The Assessment of Reinforcing
the behavior of specimens with bottom- or top-cast square Bars with Inadequate Anchorage,” Magazine of Concrete Research, V. 47,
bars or specimens with top-cast round bars; No. 171, June 1995, pp. 95-102.
3. Hassanain, M. A., and Loov, R. E., “Cost Optimization of Concrete
• Maximum load predictions based on the CEB-FIP draft Bridge Infrastructure,” Canadian Journal of Civil Engineering, V. 30,
Model Code 2010 provisions for bond are overly conserva- No. 5, Oct. 2003, pp. 841-849.
tive for all combinations of bar shapes and casting positions; 4. Mirza, M. S., and Haider, M., The State of Infrastructure in Canada:
Implications for Infrastructure Planning and Policy, McGill University,
• Square bars appear to be less sensitive to casting posi- Montreal, QC, Canada, 2003, 53 pp.
tion than round bars. Top cast factors of 0.3 and 0.6 for 5. Abrams, D. A., “Tests of Bond Between Concrete and Steel,” Univer-
round and square bars, respectively, reasonably capture sity of Illinois Bulletin No. 71, University of Illinois at Urbana-Champaign,
Urbana, IL, 1913, 240 pp.
the reductions in bond resistance based on the range of 6. Feldman, L. R., and Bartlett, F. M., “Design of a Testing Program for
parameters evaluated in this study; and Bond of Plain Reinforcement,” 5th International PhD Symposium in Civil
• A regression analysis of the specimens shows that a Engineering, Delft, the Netherlands, 2004, pp. 145-153.
7. Hassan, M. N., and Feldman, L. R., “Behavior of Lap-Spliced Plain
linear and proportional relationship for maximum load Steel Bars,” ACI Structural Journal, V. 109, No. 2, Mar.-Apr. 2012,
as a function of lap splice length, casting position, and pp. 235-243.
equivalent diameter provides a best fit for the test data. 8. Feldman, L. R., and Bartlett, F. M., “Bond Strength Variability in
Pullout Specimens with Plain Reinforcement,” ACI Structural Journal,
V. 102, No. 6, Nov.-Dec. 2005, pp. 860-867.
ACKNOWLEDGMENTS 9. Idun, E. K., and Darwin, D., “Improving the Development Character-
Financial support was provided by a Natural Science and Engineering istics of Steel Reinforcing Bars,” SM Report No. 41, University of Kansas
Council of Canada Discovery Grant for the second author and by scholar- Center for Research, Lawrence, KS, 1995, 267 pp.
ship support for the first author from the University of Saskatchewan. 10. CAN/CSA-A23, 1/A23.2-09, “Concrete Materials and Concrete
Construction/Test Methods and Standard Practices for Concrete,” Canadian
Standards Association, Mississauga, ON, Canada, 2009, 582 pp.
AUTHOR BIOS 11. ASTM A370-97a, “Standard Test Methods and Definitions for
Montserrat Sekulovic MacLean is a Junior Engineer with KTA Structural
Mechanical Testing of Steel Products,” ASTM International, West Consho-
Engineers Ltd., Calgary, AB, Canada. She received her MSc in the Depart-
hocken, PA, 1997, 52 pp.
ment of Civil and Geological Engineering at the University of Saskatch-
12. Rao, N. R. M.; Lohrmann, M.; and Tall, L., “Effects of Strain Rate
ewan, Saskatoon, SK, Canada.
on the Yield Stress of Structural Steels,” ASTM Journal of Materials, V. 1,
No. 1, May 1966, pp. 241-262.
Lisa R. Feldman, FACI, is an Associate Professor in the Department 13. Mitutoyo, “SJ-201 Surface Roughness Tester User’s Manual No.
of Civil and Geological Engineering at the University of Saskatchewan. 99MBB0796A,” Mitutoyo Corporation, Kanagawa, Japan, 2006, 190 pp.
She is a member of ACI Subcommittee 318-R, Code Reorganization, and 14. CEB-FIP, “CEB-FIP Model Code (1990),” Comité Euro-Internatio-
Chair of Joint ACI-ASCE Committee 408, Development and Splicing of nale du Béton (CEB), Thomas Telford Ltd., London, UK, 1993, 437 pp.
Deformed Bars. 15. CEB-FIP, “fib Bulletin 55: Model Code 2010, First Complete Draft—
Volume 1,” Comité Euro-Internationale du Béton (CEB), International Feder-
NOTATION ation for Structural Concrete (fib), Lausanne, Switzerland, 2010, 292 pp.
a = shear span 16. ACI Committee 318, “Building Code Requirements for Structural
d = effective depth of reinforced splice specimens Concrete (ACI 318-11) and Commentary,” American Concrete Institute,
db = diameter for round bars or side face dimension for Farmington Hills, MI, 2011, 503 pp.
square bars 17. ASTM A305-49, “Minimum Requirements for the Deformations of
db,EQ = equivalent diameter for square bars Deformed Bars for Concrete Reinforcement,” ASTM International, West
E s = modulus of elasticity of reinforcement Conshohocken, PA, 1949, 3 pp.
fc′ = concrete compressive strength 18. Howell, D. A., and Higgins, C., “Bond and Development of
fck = characteristic value of cylinder compressive Deformed Square Reinforcing Bars,” ACI Structural Journal, V. 104,
strength of concrete No. 3, May-June 2007, pp. 333-343.
fctd = design value of concrete tensile strength 19. ACI Committee 318, “Building Code Requirements for Structural
fu = ultimate strength of reinforcement Concrete (ACI 318-05) and Commentary (ACI 318R-05),” American
fyd = dynamic yield strength of reinforcement Concrete Institute, Farmington Hills, MI, 2005, 430 pp.
fys = static yield strength of reinforcement 20. CAN/CSA-A23, 3-04, “Design of Concrete Structures,” Canadian
kb = indicator variable for bar shape Standards Association, Mississauga, ON, Canada, 2004, 258 pp.
kc = indicator variable for casting position 21. Chana, P. A., “A Test Method to Establish a Realistic Bond Stress,”
Ls = spliced length of longitudinal reinforcing bars Magazine of Concrete Research, V. 24, No. 151, 1990, pp. 83-90.
P = applied load 22. Feldman, L. R., and Bartlett, F. M., “Bond Stresses Along Plain Steel
Pmax = maximum applied load Reinforcing Bars in Pullout Specimens,” ACI Structural Journal, V. 104,
(Pmax)CEB,1990 = maximum applied load predicted using CEB-FIP No. 6, Nov.-Dec. 2007, pp. 685-692.
Model Code 1990 provisions 23. Söylev, T. A., “The Effect of Fibres on the Variation of Bond Between
(Pmax)CEB,2010 = maximum applied load predicted using CEB-FIP Steel Reinforcement and Concrete with Casting Position,” Construction &
draft Model Code 2010 provisions Building Materials, V. 25, No. 4, Apr. 2011, pp. 1736-1746.
R y = bar surface roughness 24. Barnes, R. W.; Burns, N. H.; and Kreger, M. E., “Development
uave = average bond stress Length of 0.6-Inch Prestressing Strands in Standard I-Shaped Pretensioned
γc = partial factor for concrete compressive strength Concrete Beams,” Research Report 1388-1, Center for Transportation
η1 = factor to describe reinforcing type Research, University of Texas at Austin, Austin, TX, 2000, 338 pp.

330 ACI Structural Journal/March-April 2014


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 111-S29

Performance of Glass Fiber-Reinforced Polymer-Doweled


Jointed Plain Concrete Pavement under Static and
Cyclic Loadings
by Brahim Benmokrane, Ehab A. Ahmed, Mathieu Montaigu, and Denis Thebeau
Glass fiber-reinforced polymer (GFRP) dowel bars are a Joint effectiveness E based on the measured deflections of
non-corrodible and maintenance-free alternative that will poten- loaded and unloaded sides of the joint is given by Eq. (1)
tially reduce the life-cycle cost of jointed plain-concrete pavement
(JPCP), especially in harsh environmental conditions. This paper E = 2 ⋅ δ u (δ l + δ u ) × 100 (1)
investigates the performance of GFRP dowels in JPCP under
static and cyclic loads. In addition, it compares their behavior
with that of commonly used epoxy-coated steel dowels. GFRP and
where E is the joint effectiveness; δu is the deflection of the
epoxy-coated steel dowels were employed in fabricating a total unloaded slab; and δl is the deflection of the loaded slab.
of six JPCP prototypes (slab-joint). The test prototypes measured When the deflections of the slabs on both sides of the joint
2440 mm long x 610 mm wide x 254 mm deep (96 x 24 x 10 in.). The are equal, the joint is considered 100% effective. On the
slabs were cast with a butted joint; each test prototype contained other hand, if the unloaded side of the joint experiences
two dowel bars. The test parameters included: 1) dowel-bar type no deflection, the joint is considered 0% effective. ACPA
(GFRP and epoxy-coated steel); 2) dowel-bar diameter (34.9 and (1991) recommends an adequate effectiveness of 75% in
38.1 mm [1.38 and 1.50 in.] for GFRP; 28.6 mm [1.13 in.] for pavement joints.
epoxy-coated steel); and 3) loading scheme (static and cyclic). Load-transfer efficiency LTE (AASHTO 1993) is another
The test results revealed that both 34.9 and 38.1 mm (1.38 and quantitative measure for assessing joint efficiency. Equa-
1.50 in.) GFRP dowels showed crack patterns and failure modes
tion (2) provides deflection-based LTE
similar to those of the epoxy-coated steel dowels. The 34.9 and
38.1 mm (1.38 and 1.50 in.) GFRP dowels and 28.6 mm (1.13 in.)
epoxy-coated steel dowels were not affected by 1,000,000 cycles LTE = δ u δ l × 100 (2)
between 10 and 50 kN (2.25 and 11.24 kip). In addition, both the
34.9 and 38.1 mm (1.38 and 1.50 in.) GFRP dowels showed higher ACPA (1991) considers corrosion of steel dowels as one
joint effectiveness than that of the 28.6 mm (1.13 in.) epoxy-coated of the main reasons of premature concrete failure, resulting
steel dowels. This paper also discusses the results of a field appli- from corrosion initiated by the chloride from deicing salts
cation in which the GFRP dowels were implemented in a new infiltrating the joints. Because the smooth dowel bars are
concrete-pavement highway in Mirabel, QC, Canada. assumed to be frictionless and permit free relative movement
of slabs due to temperature changes, the corrosion causes
Keywords: cyclic; design; dowel; fiber-reinforced polymers; field applica-
tion; joint; joint effectiveness; pavement; static. the steel surface to exfoliate, so that the dowels lock in the
joint. Locking, in turn, exerts excessive tensile loads on the
INTRODUCTION surrounding concrete with attendant stress concentration on
Jointed plain-concrete pavements (JPCPs) are commonly the concrete and greater movement around the joint, which
constructed with contraction joints to accommodate slab hasten the rate of joint failure (FRP Dowel Bar Team 1998).
movements due to temperature and moisture variations Furthermore, corrosion of the steel dowels may restrain free
(Ioannides and Korovesis 1992). The joints may either be movement under temperature variations. The restriction of
longitudinal, parallel to traffic, or transverse, perpendicular free movement of slabs with respect to the dowels produces
to traffic. Transverse joints are placed at regular intervals, high tensile stresses, which, in turn, result in mid-slab
creating discontinuities in the pavement and forming a series cracking (William and Shoukry 2001).
of slabs. Load transfer within a series of concrete slabs takes A common method to make steel dowels more durable
place across these joints. Thus, an effective load-transfer is coating them with fusion-bonded epoxy. Unfortu-
device should be present to transfer the loads between adja- nately, epoxy-coated steel dowels exhibit the same perfor-
cent slabs (Porter 2005). mance problems normally associated with surface coating,
Dowel bars are installed at the transverse joints of the including voids and damage during transportation and
concrete slabs to reduce the deflections and stresses at handling. The concentrated corrosive mechanisms at defects
the joints while transferring the traffic load from one slab have led, in some cases, to more rapid failure than the same
to the adjacent slab (Westergaard 1928). In addition, the
ACI Structural Journal, V. 111, No. 2, March-April 2014.
dowel-bar system works well with both narrow and wide MS No. S-2012-098, doi:10.14359.51686525, was received March 16, 2012, and
reviewed under Institute publication policies. Copyright © 2014, American Concrete
joints (Maitra et al. 2009). The load-transfer efficiency of Institute. All rights reserved, including the making of copies unless permission is
a joint is assessed by joint effectiveness E, as specified by obtained from the copyright proprietors. Pertinent discussion including author’s
closure, if any, will be published ten months from this journal’s date if the discussion
the American Concrete Pavement Association (ACPA 1991). is received within four months of the paper’s print publication.

ACI Structural Journal/March-April 2014 331


steel reinforcement without an epoxy coating (FRP Dowel
Bar Team 1998).
Glass fiber-reinforced polymer (GFRP) dowels do
not corrode, and are maintenance-free. Consequently,
employing GFRP dowels in JPCP as an alternative to the
commonly used epoxy-coated dowels will eliminate the
potential of corrosion and extend the service life of jointed
concrete pavements. A few studies have been conducted to
investigate the feasibility of GFRP dowel bars as an alter-
native to steel in jointed concrete pavements (Porter et al.
1996; Davis and Porter 1998; Porter et al. 2001; Eddie et
al. 2001; Smith 2001). These investigations supported the
use of GFRP dowels as a potential solution to the corrosion Fig. 1—Details of test prototypes and GFRP dowels.
issues of steel dowels in JPCP.
Unlike the United States, which has the Federal Highway EXPERIMENTAL PROGRAM
Administration (FHWA), Canada has no single agency
responsible for funding pavement construction and rehabil- GFRP dowel characterization
itation or for setting pavement design standards. Pavement The research involved GFRP dowels measuring 34.9 and
design for Canada’s primary highway network comes under 38.1 mm (1.38 and 1.50 in.) in diameter (Fig. 1) and epoxy-
provincial jurisdiction, while the federal government retains coated steel reference dowels 28.6 mm (1.13 in.) in diam-
responsibility for national-park roadways. Each agency is eter. The GFRP dowels were made of continuous E-glass
free to use whatever design procedure it chooses for JPCP fibers in vinylester resin using the pultrusion process.
design and rehabilitation (Canadian Strategic Highway The fiber content was 80.6% by weight. The physical and
Research Program 2002). The Ministry of Transportation mechanical properties of the GFRP dowels were deter-
of Quebec (MTQ) has attempted to overcome corrosion-as- mined using the appropriate test methods provided in
sociated problems when steel dowels are used. More than CSA S807 (Canadian Standards Association 2010) and
350 km (217.5 miles) of JPCP has been built in Quebec since ACI 440.6 (ACI Committee 440 2008). Mechanical charac-
1994. The method commonly employed for more durable terization included testing representative specimens of the
pavement has been using epoxy-coated steel dowels. These GFRP dowel bars to determine transverse shear strength
dowels, however, have evidenced performance problems (ACI 440.3R [ACI Committee 440 2004], Test Method B.4),
normally associated with surface coating, including voids, interlaminate shear (short-beam test) (ASTM D4475), flex-
corrosion, and damage during transportation and handling. ural strength, and flexural modulus of elasticity (stiffness)
Through an extended collaborative project between the (ASTM D4476). Table 1 presents the physical and mechan-
MTQ and the University of Sherbrooke, new GFRP dowels ical properties of the GFRP dowels obtained from testing.
were developed, and their long-term durability performance
was assessed (Montaigu et al. 2013). In this investigation, Test prototypes
the GFRP dowels were conditioned in harsh environments A total of six JPCP prototypes were constructed and tested.
at high temperatures for different duration periods. The find- The prototypes included two reinforced with GFRP dowels
ings of this investigation demonstrated the high stability of 34.9 mm (1.38 in.) in diameter, two reinforced with GFRP
vinylester-based GFRP dowels in concrete environment. dowels 38.1 mm (1.50 in.) in diameter, and two reinforced
This paper, however, investigated the performance of these with epoxy-coated steel dowels 28.6 mm (1.13 in.) in diam-
newly developed vinylester-based GFRP dowels in JPCP eter. For each dowel type and diameter, one prototype was
under static and cyclic loadings in laboratory conditions. tested under monotonic load (Phase I), while the second one
In addition, it compared their structural performance with was tested using a cyclic-load scheme (Phase II). The test
that of epoxy-coated steel dowels in JPCP. This paper also prototypes were 2440 mm long x 610 mm wide x 254 mm
discusses the results of a field application in which the GFRP deep (96 x 24 x 10 in.). The slabs were cast with a 19 mm
dowels tested were installed in a new concrete-pavement (0.75 in.) butt joint, and each test prototype had two dowel
highway (Hwy 15 North) in Mirabel, QC, Canada. bars spaced 305 (12 in.) mm apart. Figure 1 shows the
dimensions and details of the pavement prototypes.
RESEARCH SIGNIFICANCE The length of the specimens was selected based on
GFRP dowels appeared as a possible cost-effective solu- finite-element modeling (Maitra et al. 2009), which simu-
tion for JPCP. Few studies have investigated the feasibility lated the experimental study conducted by the U.S. Naval
of using GFRP dowels in pavements. The reported study Civil Engineering Research and Evaluation Laboratory
investigated the structural performance and effectiveness (Keeton and Bishop 1957). This study revealed that the
of the newly developed vinylester-based GFRP dowel bars vertical shear force in a dowel beyond a distance of approxi-
(34.9 and 38.1 mm [1.38 and 1.50 in.]) under static and mately 1200 mm (4 ft) from the center of the load was insig-
cyclic loadings, and in a field application. In addition, it nificant. Thus, 2440 mm (8 ft) was selected as the total length
compared the behavior of JPCP with either GFRP or epoxy- for the jointed slab prototypes for this study. The geometry
coated steel dowels in transverse joints.

332 ACI Structural Journal/March-April 2014


Table 1—Physical and mechanical properties of GFRP dowel bars
Physical properties Mechanical properties
GFRP dowel diameter, mm 34.9 38.1 GFRP dowel diameter, mm 34.9 38.1
Fiber type Glass E-type Transverse shear strength, MPa 184 ± 2 173 ± 3
Resin type Vinyl ester resin Short beam shear strength, MPa 61 ± 0 54 ± 2
Fiber content, % 80.7 80.6 Four-point flexural strength, MPa 1210 ± 50 1077 ± 61
Cure ratio, % 100 100 Flexural modulus of elasticity, GPa 50.3 ± 0.5 51.6 ± 0.8
Tg, oC 124 123 — — —
Moisture uptake, % 0.06 0.07 — — —

Notes: 1 mm = 0.0394 in.; 1 MPa = 0.145 ksi; 1 GPa = 145 ksi; °C = 5/9(°F – 32).

and dimensions of the slab prototypes were consistent with Table 2—Loading schemes of test prototypes
jointed slab prototypes tested elsewhere (Eddie et al. 2001). Phase Number and prototypes Loading scheme
The prototypes were fabricated with a MTQ Type III-A
One 34.9 mm (1.38 in.)
concrete with a target 28-day compressive strength of
GFRP
35 MPa (5.1 ksi), as specified in Standard 3101 for MTQ Monotonic to 200 kN (45 kip),
normal-density mass concrete (MTQ 2009a). The concrete I One 34.9 mm (1.50 in.)
unloading, monotonic reloading
Static GFRP
mixture contained 380 kg/m3 (23.7 lb/ft3) of GUb-SF to failure.
cement, and had a water-cement ratio (w/c) of 0.42, with One 28.6 mm (1.13 in.)
high-range water-reducing admixture to maintain a mixture epoxy-coated steel
slump of 120 ± 30 mm (4.7 ± 1.18 in.) (MTQ 2009a). The One 34.9 mm (1.38 in.)
pavement prototypes were cast at the structural laboratory GFRP
One million cycles between 10
using three concrete batches. The average concrete strength II One 34.9 mm (1.50 in.) and 50 kN (2.24 to 11.24 kip) at
of the concrete batches was 48.0 ± 3.5 MPa (7.0 ± 0.5 ksi) Cyclic GFRP 15 Hz. Thereafter, static testing
based on testing of three concrete cylinders (150 x 300 mm One 28.6 mm (1.13 in.)
until failure.
[5.9 x 11.8 in.]) from each batch. epoxy-coated steel

Subgrade base layer Table 2 summarizes the loading schemes, while Fig. 2
The granular base consisted of three 100 mm (4 in.) thick shows the test setup.
layers of limestone aggregate compacted using a 90 kg For static testing (Phase I), the monotonic load was applied
(198 lb) vibrating plate. The granular mixture was prepared with a stroke-controlled rate of 0.01 mm/sec (0.02 in./min)
according to AASHTO specifications (Class A). The gran- to allow for progressive contact and loading. The load was
ular subgrade mixture consisted of 50% sand (0 to 5 mm applied using a 1000 kN (225 kip) hydraulic actuator on one
[0 to 0.2 in.]), 20% 10 mm (0.4 in.) crushed rock (5 to 14 mm side of the joint over a loading plate of 306 mm (12 in.) in
[0.2 to 0.6 in.]), and 30% 20 mm (0.8 in.) crushed rock (14 to diameter. The prototypes were loaded up to 200 kN (45 kip),
28 mm [0.6 to 1.1 in.]). Aggregates were dampened before then the load was released. Thereafter, the prototypes were
placing to maximize the compaction efficiency. Once the loaded again at the same rate until failure. E and LTE were
base was completed, a thin layer of sand was applied to the calculated at an applied load of 40 kN (9 kip) (service load,
final surface to provide contact between the concrete surface which is equal to one half the equivalent axle load) from
and subgrade. The base was extended by 300 mm (12 in.) on the deflection measurements of two linear variable differ-
all sides to allow for load distribution and prevent failure of ential transducers (LVDTs) on both joint sides (loaded
the base-layer container. The overall dimensions of the base and unloaded).
layer were 1.52 m wide x 3.35 m long x 0.30 m deep (5 x 11 x For the fatigue testing (Phase II), the prototypes were
1 ft). Upon completing the base, the base modulus (stiffness) tested up to 1 million cycles. The load followed a sinusoidal
was measured using a Briaude Compacting Device (BCD), waveform that varied from 10 to 50 kN (2.25 to 11.24 kip).
and was 110 MPa/m (4.9 ksi/ft). The minimum load (10 kN [2.25 kip]) was required to main-
tain contact between the slab and loading plate and to mini-
Testing loads and procedures mize the impact on the subgrade. The maximum load (50 kN
The JPCP prototypes were tested under two different [11.24 kip]) was set to achieve the service load and keep
loading conditions: static (Phase I) and cyclic (Phase II). 40 kN (9 kip) as the cyclic test amplitude, which is equal to
During Phase I, the prototypes were monotonically loaded one-half the equivalent axle load (service load). It should
to 200 kN (45 kip) to induce cracks at the joints. Thereafter, be mentioned that this loading scheme closely represents
the load was released, and the prototypes were loaded again field conditions under which load is applied and removed as
up to failure. During Phase II, the prototypes were subjected a vehicle approaches the joint or moves away from it. The
to 1 million cycles ranging from 10 to 50 kN (2.25 and load was applied with the same hydraulic actuator (1000 kN
11.24 kip), followed by monotonic loading up to failure. [225 kip]) with a load-controlled scheme. The loading

ACI Structural Journal/March-April 2014 333


Fig. 2—Test setup: (a) overall view; and (b) loading plate and linear variable displacement transducers.
Table 3—Summary of test results
Cracking and failure loads of jointed pavement prototypes
Phase Load at Steel, 28.6 mm GFRP, 34.9 mm GFRP, 38.1 mm

I Cracking, kN (kip) 140.7 (31.6) 100.0 (22.5) 124.8 (28.1)


Static Failure, kN (kip) 506.6 (113.9) 460.0 (103.4) 478.0 (107.5)

II Cracking, kN (kip) 250 (56.2) 178 (40.0) 145 (32.6)


Cyclic Failure, kN (kip) 622 (139.8) 526 (118.2) 413 (92.8)
Joint effectiveness and load-transfer efficiency at service load, 40 kN (9 kip)
Steel, 28.6 mm GFRP, 34.9 mm GFRP, 38.1 mm
Phase Load type E, % LTE, % E, % LTE, % E, %

I First loading 86 75 89 81 95
Static Reloading 65 45 64 47 74

II First loading 95 90 95 90 96
Cyclic After cycling 92 85 93 87 95

Note: 1 mm = 0.0394 in.

and unloading was applied at a frequency of 15 Hz. This Phase I: 28.6 mm (1.13 in.) steel dowels, 34.9 mm (1.38 in.)
frequency is equivalent to the time that a vehicle needs to GFRP dowels, and 38.1 mm (1.50 in.) GFRP dowels.
cross the joint, assuming a speed of 65 to 80 kph (37.3 to In the case of the 28.6 mm (1.13 in.) diameter steel dowels,
49.7 mph) (MTQ 2009b). Because the test prototypes did the first crack appeared in the loaded slab at 140.7 kN
not fail after 1 million cycles, the prototypes were retested (31.6 kip) on one side, and at 197.0 kN (44.3 kip) on the
under monotonic static load until failure. Before cycling, as other. The crack appeared at the level of dowel bars, and
well as after predetermined sets of cycles (1; 1000; 10,000; continued to the surface. The second crack appeared under a
100,000; 500,000; and 1,000,000), the load cycling was load of approximately 380 kN (85.4 kip). The slab prototype
interrupted, and a monotonic loading test up to 40 kN (9 kip) failed at 506.6 kN (113.9 kip) by shear failure of the loaded
(service load) was conducted to assess joint performance. slab beyond the dowel bars. The GFRP-doweled prototypes
showed crack patterns and failure modes similar to the
TEST RESULTS steel-doweled prototype. The cracking loads were 100 and
124.8 kN (22.4 and 28.1 kip) for both 34.9 and 38.1 mm
Static testing (Phase I) (1.38 and 1.50 in.) diameter GFRP dowels, respectively.
Cracking and failure—When the JPCP prototypes were These loads represent 71 and 89% of the cracking load of
submitted to 200 kN (45 kip), cracks appeared in the loaded the steel-doweled prototype. On the other hand, the failure
slabs. The unloaded slabs, however, did not evidence any loads for the 34.9 and 38.1 mm (1.38 and 1.50 in.) diameter
cracks during the test. Table 3 gives the cracking and failure GFRP dowels were 460 and 478 kN (103.4 and 107.4 kip),
loads of the test prototypes. Figure 3 shows the cracking respectively, which represent 91 and 94% of the failure load
patterns and failure modes of the three prototypes tested in of the steel-doweled prototype.

334 ACI Structural Journal/March-April 2014


Fig. 3—Cracking at failure of Phase I prototypes: (a) 28.6 mm steel dowels; (b) 34.9 mm GFRP dowels; and (c) 38.1 mm GFRP
dowels. (Note: 1 mm = 0.0394 in.)

Fig. 4—Results of Phase I prototypes (static testing up to 200 kN): (a) joint effectiveness; (b) load-transfer efficiency; and
(c) relative deflection. (Note: 1 mm = 0.0394 in.; 1 kN = 0.225 kip.)
Joint effectiveness E and load-transfer efficiency LTE— ence between the two prototypes with 35.9 mm (1.38 in.)
E and LTE were calculated using the deflection measure- GFRP dowels and 28.6 mm (1.13 in.) steel dowels after
ments recorded by the two LVDTs placed on the unloaded 50 kN (11.24 kip). The differences between the two proto-
and loaded sides of the joint. Table 3 gives the calculated E types under 50 kN (11.24 kip) may be related to the better
and LTE at service load (40 kN [9 kip]) for the tested JPCP compaction of the subgrade base after testing the first proto-
prototypes with steel and GFRP dowels. All tested proto- type, which had 28.6 mm (1.13 in.) steel dowels. Further-
types showed E and LTE higher than 75 and 60%, respec- more, increasing the GFRP dowels to 38.1 mm (1.50 in.)
tively, which meets ACPA (1991) requirements. increased E and LTE. Figure 4(a) also shows that cracking
Both GFRP dowel diameters (34.9 and 38.1 mm [1.38 and during the first loading did not affect E, because the load was
1.50 in.]) displayed E and LTE higher than the 28.6 mm not released until it reached 200 kN (45 kip).
diameter (1.13 in.) steel dowels. The 34.9 mm (1.38 in.) Relative deflection—Figure 4(c) provides the relative
diameter GFRP dowels showed E of 89% and LTE of 81%, deflection of loaded and unloaded slabs of three prototypes
while the 38.1 mm (1.50 in.) diameter GFRP dowels showed with 28.6 mm (1.13 in.) epoxy-coated steel dowels, 34.9 mm
E of 95% and LTE of 92%. The values in Table 3 reveal that (1.38 in.) GFRP dowels, and 38.1 mm (1.50 in.) GFRP
using 34.9 mm (1.38 in.) diameter GFRP dowels instead of dowels. The relative deflection of the 28.6 mm (1.13 in.) steel
28.6 mm diameter (1.13 in.) steel dowels increased E and dowels was between those of the 34.9 and 38.1 mm (1.38 and
LTE by 9 and 8%, respectively. On the other hand, replacing 1.50 in.) GFRP dowels. Figure 4(c) shows immediate deflec-
28.6 mm diameter (1.13 in.) steel dowels with 38.1 mm tion at the beginning of the test ranging from 0.2 to 0.45 mm
diameter (1.50 in.) GFRP dowels increased E and LTE by (0.01 to 0.02 in.). This immediate deflection occurred because
10 and 23%, respectively. In addition, E and LTE values the specimens were not cast directly on the subgrade base
revealed that the behavior of the jointed pavement with and, when the load was applied to the pavement prototypes,
34.9 mm (1.38 in.) diameter GFRP dowels was almost the the immediate deflection occurred until complete contact
same as that with 28.6 mm (1.13 in.) diameter steel dowels. between the concrete surface and the subgrade base layer was
Furthermore, Table 3 shows that reloading the specimens achieved. The steel-doweled pavement prototype showed the
after cracking during the initial loading phase (200 kN highest immediate deflection (≈0.45 mm [0.02 in.]) because
[45 kip]) yielded very low E and LTE because of the cracks. it was the first tested, and may have been affected by the
The 38.1 mm (1.50 in.) diameter GFRP dowels evidenced compressibility of the subgrade layer. The immediate deflec-
joint effectiveness E of 74%, which is very close to 75%, as tion was less in the case of the GFRP-doweled pavement
provided for by ACPA (1991). Therefore, JPCP stability and prototypes (≈0.20 mm [0.01 in.]). This immediate deflection
performance is dependent on the slabs remaining uncracked increase might not occur in field applications in which the
to achieve efficient joints. pavement is cast directly on the subgrade.
E and LTE were plotted against applied load in Fig. 4(a) At service load (40 kN [9 kip]), the 28.6 mm (1.13 in.)
and (b). It should be mentioned that LTE corresponding to steel dowels showed a relative deflection of 0.58 mm
E = 75% was 60%. Figure 4(a) and (b) demonstrate that, (0.02 in.). The 34.9 and 38.1 mm (1.38 and 1.50 in.) GFRP
after an initial loading interval till about 50 kN (11.24 kip), E dowels evidenced relative deflections of 0.82 and 0.36 mm
and LTE stabilized. Besides, there was no significant differ- (0.03 and 0.01 in.), respectively. The relative deflections

ACI Structural Journal/March-April 2014 335


Fig. 5—Cracking at failure of Phase II prototypes: (a) 28.6 mm steel dowels; (b) 34.9 mm GFRP dowels; and (c) 38.1 mm
GFRP dowels. (Note: 1 mm = 0.0394 in.)
of the 34.9 and 38.1 mm (1.38 and 1.50 in.) GFRP dowels (1.38 and 1.50 in.) GFRP dowels showed failure loads
were 1.41 and 0.62 times that of the 28.6 mm (1.13 in.) steel equal to 85 and 66% of the failure load of the steel-doweled
dowels, respectively. After the service load, the difference prototype. Figure 5 shows the cracking patterns and failure
between the relative deflections of the 28.6 mm (1.13 in.) modes of the three prototypes tested in Phase II: 28.6 mm
steel dowel and 34.9 mm (1.38 in.) GFRP dowels decreased. (1.13 in.) steel dowels, 34.9 mm (1.38 in.) GFRP dowels,
At approximately 70 kN (15.7 kip), the difference between and 38.1 mm (1.50 in.) GFRP dowels. In addition, Table 3
the two prototypes was less than 0.1 mm (0.004 in.). On the also shows that the prototype with 38.1 mm (1.50 in.)
other hand, the relative deflection of the 38.1 mm (1.50 in.) GFRP dowels yielded lower cracking load and failure load
GFRP dowels was very small compared with that of the compared with that of the prototype with 34.9 mm (1.38 in.)
28.6 mm (1.13 in.) epoxy-coated steel dowels. GFRP dowels. The normal variance in the concrete strength
between the different batches may have had an effect on the
Cyclic testing (Phase II) lower cracking and failure loads.
Cracking and failure—The three prototypes with steel Joint effectiveness E and load-transfer efficiency LTE—
and GFRP dowels did not experience any cracking after Before cycling, as well as after predetermined sets of cycles
1,000,000 cycles at 40 kN (9 kip) (between 10 and 50 kN (1; 1000; 10,000; 100,000; 500,000; and 1,000,000), the
[2.25 and 11.24 kip]). Thus, using durable dowel bars in load cycling was interrupted, and a monotonic loading test
jointed pavements will yield efficient joints with extended up to 40 kN (9 kip) (service load) was conducted to assess
service life under service load. This confirms the findings joint performance. Table 3 lists the calculated E and LTE at
of the static testing (Phase I): JPCP efficiency will not be service load (40 kN [9 kip]) for the JPCP prototypes with
altered as long as the concrete does not crack. steel and GFRP dowels. All tested prototypes showed E and
After 1,000,000 cycles, the three prototypes were retested LTE higher than 75 and 60%, respectively, which meets
under monotonic load up to failure. Table 3 lists the cracking ACPA (1991) requirements. Table 3 shows that E and LTE
and failure loads. The cracking load of the prototypes in in Phase II were higher than in Phase I. Once again, this is
Phase II was higher than those in the Phase I prototypes. The related to the excessive compaction resulting from the cyclic
cyclic testing of the three prototypes affected the subgrade testing of the prototypes. The continued cycling resulted
base, and resulted in very high compaction. That, in turn, in further compaction of the base layer, which increased
affected the cracking loads of the test prototypes. Similarly, joint effectiveness.
the failure loads in Phase II were also higher than those in E and LTE were plotted against applied load for a selected
Phase I, except for the prototype with 38.1 mm (1.50 in.) number of cycles in Fig. 6. The 38.1 mm (1.50 in.) GFRP
GFRP dowels, which showed 413 kN (92.9 kip) compared dowels showed the highest E and LTE, which were 96 and
with 478 kN (107.4 kip) in Phase I (static testing). 92%, respectively. Figure 6 also reveals no significant
Table 3 shows that the prototype with 28.6 mm (1.13 in.) difference in E and LTE after 1,000,000 cycles. There was
epoxy-coated dowels experienced the highest cracking no significant difference between the two prototypes with
load among the tested prototypes. The prototypes with 34.9 mm (1.38 in.) GFRP dowels and 28.6 mm (1.13 in.)
34.9 and 38.1 mm (1.38 and 1.50 in.) GFRP dowels had steel dowels (E and LTE equal 95 and 90%, respectively).
cracking loads of 71 and 58% of that of the prototype with
steel dowels. As with Phase I prototypes, the first cracks DESIGN OF GFRP DOWELS
appeared at the joints at the level of the dowel bars, and then Because the properties of GFRP dowel are different from
extended to the surface. Compared with Phase I, Phase II those of steel, directly replacing steel dowels with GFRP
prototypes showed more cracks at failure, except in the case dowels of the same diameter and at the same spacing is
of the 38.1 mm (1.50 in.) GFRP-doweled prototype, which not valid. The required diameter and spacing can be deter-
showed one major crack. The cracks led to the splitting of mined by equating the relative deflection of a joint doweled
concrete around the dowels before failure, which occurred with steel to that of a joint doweled with GFRP (Davis and
due to shear failure of the loaded slabs past dowel length. Porter 1998). The effective design steps for GFRP-doweled
The 28.6 mm (1.13 in.) steel-doweled prototype failed at JPCP are: 1) determine the load transferred by the critical
622 kN (139.8 kip). The prototypes with 34.9 and 38.1 mm dowel; 2) determine the relative deflection for a joint with

336 ACI Structural Journal/March-April 2014


Fig. 6—Results of Phase II prototypes (cyclic loading) at 40 kN: (a) joint effectiveness; (b) load-transfer efficiency. (Note:
1 mm = 0.0394 in; 1 kN = 0.225 kip.)
steel dowels; 3) determine the relative deflection for a joint
with GFRP dowels; 4) determine the required diameter and
spacing for GFRP dowels; and 5) check the bearing stresses.
This design procedure is summarized as follows.
When a load is applied to the edge of a slab, a portion of the
load is transferred to the adjacent slab through the dowels by
shear. Tabatabie et al. (1979) suggested that only the dowels
located within a distance of lr from the load point contributes
to transferring the load (Fig. 7(a)), where lr is the radius of
relative stiffness as defined in Eq. (3) by Westergaard (1925)

( )
lr = 4 Ec h3 12 1 − µ 2 k (3)

where Ec is the modulus of elasticity of the pavement


concrete; h is the pavement thickness; μ is the Poisson’s
ratio of the pavement concrete; and k is the modulus of
subgrade reaction.
Corresponding to the contribution of the dowels located in
the lr distance, the load transferred by the critical dowel Pt is
Fig. 7—Load-transfer distribution and relative deflection
given by Eq. (4)
between jointed slabs.
Design Load Transfer ( Pd ) According to Friberg (1938), yo is calculated as in Eq. (7).
Pt = (4)
Number of Effective Dowels This equation is derived assuming the dowel bars have a

semi-infinite length. Albertson (1992), however, showed
that this equation can be applied to dowel bars with a βL
Yoder and Witczak (1975) reported that a 5 to 10% reduc-
value greater than 2 with a minor error, where L is the length
tion in load transfer occurred due to the formation of voids
of dowel bar on one side of the slab
beneath the dowels at the joint face. Accordingly, a design
load transfer of 45% of the applied wheel load is recom-
Pt
mended. Thus, the design load transfer Pd is calculated from yo = (2 + b z ) such that βL > 2.0 (7a)
4 b3 E I
Eq. (5) as a function of the applied wheel load Pw

Pd = 0.45Pw (5) b = 4 Kb 4 EI (7b)


Considering the schematic shown in Fig. 7(b), the rela- where yo is the dowel deflection relative to or within concrete
tive deflection between the jointed slabs Δ is calculated from at the face; Pt is the load transferred by the critical dowel;
Eq. (6), neglecting the deflection due to the slope and flexure β is the relative stiffness of the dowel bar encased in the
along the joint width concrete; L is the length of dowel bar on one side of the
slab; E is the flexural modulus of elasticity; I is the moment
∆ = 2yo + δ (6) of inertia; b is the dowel diameter; and K is the modulus of
dowel support or reaction.
where yo is the dowel deflection relative to or within the It should be mentioned that K is an important param-
concrete at the face; and δ is the shear deflection of the eter in the Friberg (1938) design equation. K is determined
dowel across the joint. empirically because of the difficulty in establishing it theo-
retically (Friberg 1938). Yoder and Witczak (1975) found

ACI Structural Journal/March-April 2014 337


that K ranges between 81.43 and 407.17 N/mm3 (3 × 105 8) transverse shear strength (ACI 440.3R [ACI Committee
and 1.5 × 106 lb/in.3). For the analytical calculations herein, 440 2004], Test Method B.4) ≥ 170 MPa (24.7 ksi).
a value of 407.17 N/mm3 (1.5 × 106 lb/in.3) has been used To take a step forward toward employing GFRP dowels
as suggested to simulate the worst-case scenario (Yoder and in field applications, MTQ has implemented the use of
Wiczak 1975). GFRP dowels in a demonstration section on Hwy 15 North
The shear deflection of the dowel across the joint is then in Mirabel, QC, Canada (just to the north of HW 50). The
calculated as shown in Eq. (8) traffic volume on this highway is approximately 100,000
vehicles per day, with trucks approaching 6%. The experi-
δ = lPt z AG (8) mental section is in the right lane and the center of a length
of 500 m (1640 ft) divided into two parts: 250 m (820 ft) for
where δ is the dowel’s shear deflection across the joint; l 38.1 mm (1.50 in.) GFRP dowels, and 250 m (820 ft) for
is the shear shape factor; Pt is the load transferred by the 34.9 mm (1.38 in.) steel dowels. Figure 8 shows the GFRP
critical dowel; z is the joint width; A is the area of dowel bar; dowel arrangement during the construction of the highway.
and G is the shear modulus, which is assumed as 3.3 GPa Both the GFRP and steel dowels were 450 mm (18 in.) long
(478.6 ksi) based on the literature. and spaced at 300 mm (12 in.). The pavement thickness was
The bearing stress developed due to slab deformation 260 mm (10.2 in.), while the length of the jointed slabs was
was calculated as shown in Eq. (9), which should be less set at 5000 mm (197 in.). The joints at the dowel locations
than the permissible bearing stress underneath the pavement were sawn within a few hours of casting and allowed to crack
provided by ACI Committee 325 (1956) due to thermal contraction and shrinkage. The JPCP was cast
using MTQ Type IIIA concrete with a 28-day compressive
σ b = Kyo < fb = 1 3 ⋅ ( 4 − b ) fc′ (9) strength of 35 MPa (5.1 ksi) according to MTQ requirements
(MTQ 2009a).
where σb is the developed bearing stress; K is the dowel’s As part of an ongoing MTQ effort, the performance of
modulus of support; yo is dowel deflection as the concrete the experimental section with the GFRP and steel dowels is
face; fb is the permissible bearing stress; b is the dowel diam- being assessed with a falling-weight deflectometer (FWD)
eter; and fc′ is the concrete compressive strength. (ASTM D4694) to locate and quantify weak areas that can
Considering the geometries and material properties of the be corrected before premature damage occurs. The FWD
dowels used in this investigation, a comparable design was is a nondestructive field test that involves applying impact
conducted to determine the diameter of GFRP dowels that loads to the pavement surface and monitoring the pavement
could replace 28.6 mm (1.13 in.) diameter epoxy-coated steel deflection response through a series of velocity transducers
dowels. The design included the three diameters of GFRP placed on the pavement at specified distances from the load,
dowels used herein: 31.8, 34.9, and 38.1 mm (1.25, 1.38, as illustrated in the FWD test diagram in Fig. 9(a). To simu-
and 1.50 in.). Provided in Table 4, it shows that replacing late moving-axle loads, the FWD load is applied by releasing
28.6 mm (1.13 in.) epoxy-coated steel dowels with 31.8 mm a mass from a selected height to impact the road surface. The
(1.25 in.) GFRP dowels increased the bearing stress and rela- mass and drop height of the load determine the peak contact
tive deflection by 23 and 88%, respectively. Using 34.9 mm pressure applied to the road surface.
(1.38 in.) GFRP dowels yielded almost the same bearing The FWD test was conducted using the MTQ testing
stress as that of 28.6 mm (1.13 in.) epoxy-coated dowels, vehicle (Fig. 9(b)). The falling weight is lifted to the desired
while the relative deflection was 58% higher. On the other height and dropped on an anvil fitted with rubber shock
hand, using 38.1 mm (1.50 in.) GFRP dowels yielded 29% absorbers, inducing an impact of 40 kN (9 kip). The anvil
lower bearing stress than the 28.6 mm (1.13 in.) epoxy-coated transmits the force of impact to the pavement through a plate
steel dowels, while the relative deflection was 33% higher. placed on the road surface. A series of sensors simultane-
Considering the bearing stress as reference, the 34.9 mm ously measure pavement vertical movement resulting from
(1.38 in.) GFRP dowels may be considered as a direct alter- a shock wave. This movement is very subtle, because it
native to the 28.6 mm (1.13 in.) epoxy-coated dowels when lasts only 30 ms, and is measured in microns. The specifica-
the materials and conditions in this investigation prevail. tions of the MTQ testing apparatus are load-plate diameter:
300 mm (12 in.); falling weight: 250 kg (551 lb); impact
FIELD APPLICATION strength: 40 kN (9 kip); and location of sensors: 0 to 1.8 m
Based on the results of this research project and the charac- (0 to 5.9 ft) from the center of the plate (seven to 15 sensors).
teristics provided in Table 1, the MTQ has issued new materials In 2011, 20 FWD tests were conducted on each section
specifications for GFRP dowels as minimum requirements for (steel and GFRP). The maximum, minimum, and average
their use in JPCP in the province of Quebec (Montaigu et al. LTE for the GFRP-doweled sections were 85.2, 93.6, and
2013): 1) the fiber type should be (E-glass) or (ECR-glass); 88.6 ± 2.6%, respectively. The maximum, minimum, and
2) vinylester resin; 3) glass-fiber content (ASTM D3171, average LTE for the steel-doweled sections were 85.3, 91.6,
Procedure G) ≥ 75%; 4) moisture absorption (ASTM D570) and 87.9 ± 1.8%, respectively. In addition, core samples
≤ 0.15%; 5) glass transition temperature (ASTM D3418) ≥ were extracted from the experimental sections during the
110°C (230°F); 6) cure ratio (CSA S807) ≥ 98%; 7) short- FWD testing, yielding an average compressive strength of
beam shear strength (ASTM D4475) ≥ 59 MPa (8.6 ksi); and 49.6 MPa (7.2 ksi). These results indicate that the steel- and
GFRP-doweled slabs had almost the same performance with

338 ACI Structural Journal/March-April 2014


Table 4—Design for alternative GFRP dowel bar diameter
Reference Alternative 1 Alternative 2 Alternative 3
Design parameter steel GFRP GFRP GFRP
Dowel diameter b, mm 28.6 31.8 34.9 38.1
Dowel area A, mm2 642.42 794.23 956.62 1140.09
Moment of inertia I, mm4 32,842 50,197 72,824 103,436
Dowel modulus of elasticity E, MPa 200,000 52,600 50,300 51,600
Geometry and material properties

*
Dowel shear modulus G, MPa 78,000 3300 3300 3300
Dowel length L, mm 457 457 457 457
Dowel spacing s, mm 305 305 305 305
Concrete strength fc′, MPa 48 48 48 48
Concrete’s modulus of elasticity Ec,
32,909 32,909 32,909 32,909
MPa
Concrete’s Poisson’s ratio μ 0.2 0.2 0.2 0.2
Pavement thickness h, mm 254 254 254 254
Modulus of subgrade reaction k, MPa/m 110 110 110 110
3
Modulus of dowel support K, N/mm 407.17 407.17 407.17 407.17
Joint width z, mm 19 19 19 19

Ec h 3
lr = , mm
( ) 808 808 808 808
4
12 1 − µ 2 k

Number of dowels 1.87 1.87 1.87 1.87


Design load transfer = 0.45Pw, kN 18 18 18 18
Load transferred by critical dowel Pt 9.63 9.63 9.63 9.63

Kb
b= 4 , mm–1 0.026 0.033 0.031 0.029
4E I
Design

βL > 2.0 (ok) 5.90 7.60 7.17 6.67

Pt
yo = (2 + b z ) , mm 0.053 0.065 0.055 0.046
4 b3 E I

l Pt z
δ= , mm 0.004 0.078 0.064 0.054
AG

Δ = 2yo + δ, mm 0.110 0.208 0.175 0.147


σb = Kyo, MPa 21.63 26.51 22.47 18.84

fb =
(4 − b) fc′ , MPa 63.54 63.49 63.44 63.39
3

σb GFRP/σb steel — 1.23 1.04 0.71


Δ GFRP/Δ steel — 1.88 1.58 1.33

Note: 1 mm = 0.0394 in.; 1 kN = 0.225 kip; 1 MPa = 0.145 ksi; 1 mm2 = 0.00155 in.2; 1 N/mm3 = 0.2714 kip/in.3; 1 MPa/m = 0.0442 ksi/ft; 1 mm–1 = 25.4 in.–1

an average LTE of 87.9 ± 1.8% and 88.6 ± 2.6% for the the 75% specified by ACPA (1991). This indicates that the
steel- and GFRP-doweled slabs, respectively. The LTE of GFRP dowels tested performed efficiently in transferring
both the steel- and GFRP-doweled slabs were higher than loads in JPCP.

ACI Structural Journal/March-April 2014 339


SUMMARY AND CONCLUSIONS 38.1 mm (1.38 and 1.50 in.) GFRP dowels were 2.5 and 3.1
Through an extensive collaboration project between times the service load (40 kN [9 kip]), respectively;
the Ministry of Transportation of Quebec (MTQ) and the 3. Under the static testing in Phase I, the pavement proto-
University of Sherbrooke, new GFRP dowels were devel- type with 28.6 mm (1.13 in.) steel dowels and those with
oped, and their long-term durability characteristics were 34.9 and 38.1 mm (1.38 and 1.50 in.) GFRP dowels showed
assessed (Montaigu et al. 2013). This study, however, similar crack patterns and modes of failure. The capacities of
investigated the performance of the newly developed GFRP the three prototypes were 506.6, 460.0, and 478.0 kN (113.9,
dowels (34.9 and 38.1 mm [1.38 and 1.50 in.]) in JPCP 103.4, and 107.5 kip), respectively;
under static and cyclic loadings. In addition, it compared 4. Under the cyclic testing in Phase II, the prototypes
their structural performance with that of epoxy-coated steel resisted 1 million cycles at service load without cracking.
dowels in JPCP. Based on the test results presented and Thus, providing durable dowel bars capable of withstanding
discussed herein, the following conclusions concerning the the environmental conditions and deterioration will yield
tested GFRP dowels can be drawn: pavements with extended service lives;
1. The GFRP dowel bars provided suitable and effec- 5. Phase II produced higher cracking and ultimate loads
tive alternatives to the common epoxy-coated steel dowels compared with Phase I. The results were similar for joint
to overcome corrosion problems and related deterioration effectiveness E and load-transfer efficiency LTE. This can
under static and cyclic loading conditions; be accounted for by excessive compaction resulting from
2. Under the static testing in Phase I, the pavement proto- prototype cyclic testing, which tended to increase joint
type with 28.6 mm (1.13 in.) steel dowels showed its first effectiveness;
crack at 140.7 kN (31.6 kip), while the prototypes with 34.9 6. The pavement prototypes with GFRP dowels (34.9 and
and 38.1 mm (1.38 and 1.50 in.) GFRP dowels showed their 38.1 mm [1.38 and 1.50 in.]) showed joint effectiveness E
first cracks at 100 and 124.8 kN (22.4 and 28.1 kip), respec- and load-transfer efficiency LTE similar to or higher than
tively. The cracking loads of the prototypes with 34.9 and that of the prototypes with steel dowels (28.6 mm [1.13 in.]);
7. Achieving stability and good performance in the
GFRP-doweled JPCP requires that slabs remain uncracked
to enable achieving efficient joints and maintain efficient
load transfer;
8. Considering the tested material and setup configura-
tion, the 34.9 mm (1.38 in.) diameter GFRP dowels behaved
very similarly to the 28.6 mm (1.13 in.) diameter steel
dowels. Thus, the 34.9 mm (1.38 in.) GFRP dowels may be
viable alternatives to the 28.6 mm (1.13 in.) epoxy-coated
steel dowels subjected to the same loading and boundary
conditions. The test design also indicates that the 34.9 mm
(1.38 in.) GFRP dowels could be direct alternatives to the
28.6 mm epoxy-coated steel dowels; and
9. The field application showed similar LTE values for of
the 38.1 mm (1.50 in.) GFRP dowels and 34.9 mm (1.38 in.)
steel dowels in real service conditions. The steel- and
Fig. 8—GFRP dowel arrangement during construction GFRP-doweled slabs had average LTEs of 87.9 ± 1.8% and
(Hwy 15, Mirabel, QC, Canada).

Fig. 9—In-place falling-weight deflectometer (FWD) test (Hwy 15, Mirabel, QC, Canada): (a) schematic; and (b) in-place test.

340 ACI Structural Journal/March-April 2014


88.6 ± 2.6%, respectively, which were higher than the 75% ASTM D570, 2010, “Standard Test Method for Water Absorption of
Plastics,” ASTM International, West Conshohocken, PA, 4 pp.
specified by ACPA (1991). CAN/CSA S807-10, 2010, “Specification for Fibre-Reinforced Poly-
mers,” Canadian Standards Association (CSA), Rexdale, ON, Canada, 27 pp.
ACKNOWLEDGMENTS Canadian Strategic Highway Research Program (C-SHRP), 2002, “Pave-
The authors wish to acknowledge the financial support of Quebec’s ment Structural Design Practices Across Canada,” C-SHRP Technical Brief
Ministry of Transport (MTQ), the Natural Sciences and Engineering No. 23, Ottawa, ON, Canada, 10 pp.
Research Council of Canada (NSERC), and Quebec’s Fonds québécois de Davis, D., and Porter, M. L., 1998, “Evaluation of Glass Fiber Reinforced
la recherche sur la nature et les technologies (FQRNT). The authors are Polymer Dowels as Load Transfer Devices in Highway Pavement Slabs,”
grateful to Pultrall Inc. for providing the GFRP dowel bars and RocTest Proceedings of Transportation Conference, Ames, IA, pp. 78-81.
Inc. for providing the Briaude Compacting Device (BCD). Special thanks Eddie, D.; Shalaby, A.; and Rizkalla, S., 2001, “Glass Fiber-Reinforced
to the technical staff at the Department of Civil Engineering for their help Polymer Dowels for Concrete Pavements,” ACI Structural Journal, V. 98,
in fabricating and testing the specimens. No. 2, Mar.-Apr., pp. 201-206.
Friberg, B. F., 1938, “Load and Deflection Characteristics of Dowels in
Transverse Joints of Concrete Pavements,” Proceedings of Highway Research
AUTHOR BIOS Board No. 18, National Research Council, Washington, DC, pp. 140-154.
Brahim Benmokrane, FACI, is Tier-1 Canada Research Chair Professor FRP Dowel Bar Team (DBT), 1998, “Fiber-Reinforced Polymer (FRP)
in Advanced Composite Materials for Civil Structures and an NSERC Composite Dowel Bars—15 Years of Durability Study,” Market Develop-
Research Chair Professor in Fiber-Reinforced Polymer Reinforcement ment Alliance, Composites Institute, Harrison, NY, 18 pp.
for Concrete Infrastructure in the Department of Civil Engineering at the Ioannides, A. M., and Korovesis, G. T., 1992, “Analysis and Design of
University of Sherbrooke, Sherbrooke, QC, Canada. He is a member of ACI Doweled Slab-on-Grade Pavement Systems,” Journal of Transportation
Committee 440, Fiber-Reinforced Polymer Reinforcement. Engineering, V. 118, No. 6, pp. 745-768.
Keeton, J. R., and Bishop, J. A., 1957, “Load Transfer Characteristics of
ACI member Ehab A. Ahmed is a Postdoctoral Fellow in the Depart- a Dowelled Joint Subjected to Aircraft Wheel Loads,” Proceedings of the
ment of Civil Engineering at the University of Sherbrooke, and Lecturer at Highway Research Board, No. 36, Transportation Research Board, Wash-
Menoufiya University, Shebin El-Kom, Minoufiya, Egypt. He received his ington, DC, pp. 190-198.
PhD in civil engineering from the University of Sherbrooke. His research Maitra, S. R.; Reddy, K. S.; and Ramachandra, L. S., 2009, “Load
interests include structural analysis, design and testing, and structural Transfer Characteristics of Dowel Bar System in Jointed Concrete Pave-
health monitoring of fiber-reinforced polymer concrete structures. ment,” Journal of Transportation Engineering, V. 135, No. 11, pp. 813-821.
Ministry of Transportation of Québec (MTQ), 2009a, “Bétons de masse
Mathieu Montaigu is an MSc Student in the Department of Civil Engi- volumique normale,” Norme 3101, Ministère des Transports du Québec,
neering at the University of Sherbrooke. His research interests include Québec, Canada, 8 pp.
concrete durability and performance evaluation of glass fiber-reinforced Ministry of Transportation of Québec (MTQ), 2009b, “Bulletin Inno-
polymer dowel bars for concrete pavement. vation Transport,” No. 34, Ministère des Transports du Québec, Quebec,
Canada, Jan., 42 pp.
Denis Thebeau is a Professional Engineer with the Ministry of Trans- Montaigu, M.; Robert, M.; Ahmed, E.; and Benmokrane, B., 2013,
portation of Quebec (Pavement Division), Quebec City, QC, Canada. His “Laboratory Characterization and Evaluation of the Durability Perfor-
research interests include the design, construction, rehabilitation, and mance of New Polyester and Vinylester E-Glass GFRP Dowels for Jointed
long-term monitoring of the performance of concrete pavements. Concrete Pavement,” Journal of Composites for Construction, ASCE,
V. 17, No. 2, pp. 176-187.
Porter, M. L., 2005, “Structural Dowel Bar Alternative and Gaps of
REFERENCES Knowledge,” Proceedings of the Mid-Continent Transportation Research
AASHTO, 1993, “Guide for Design of Pavement Structures,” American
Symposium, Ames, IA, Aug., 13 pp.
Association of State Highway and Transportation Officials, Washington,
Porter, M. L.; Davis, D.; Guinn, R.; Lundy, A.; and Rohner. J., 2001,
DC, 624 pp.
“Investigation of Glass Fiber Composite Dowel Bars for Highway Pave-
ACI Committee 325, 1956, “Structural Design Considerations for Pave-
ment Slabs,” Final Report TR-408, submitted to Highway Division of the
ment Joints,” ACI Journal, V. 53, No. 7, July, pp. 1-29.
Iowa Department of Transportation and Iowa Highway Research Board,
ACI Committee 440, 2008, “Metric Specification for Carbon and Glass
Iowa State University, Engineering Research Institute, Ames, IA, 168 pp.
Fiber-Reinforced Polymer Bar Materials for Concrete Reinforcement (ACI
Porter, M. L.; Hughes, B. W.; and Barnes, B. A., 1996, “Fiber Composite
440.6M-08),” American Concrete Institute, Farmington Hills, MI, 10 pp.
Dowels in Highway Pavements,” Proceedings of the Semisequicentennial
ACI Committee 440, 2004, “Guide Test Methods for Fiber-Reinforced
Transportation Conference, May, Iowa State University, Ames, IA, 6 pp.
Polymers (FRPs) for Reinforcing or Strengthening Concrete Structures (ACI
Smith, K. D., 2001, “Alternative Dowel Bars for Load Transfer in Jointed
440.3R-04),” American Concrete Institute, Farmington Hills, MI, 39 pp.
Concrete Pavements,” FHWA Technical Report Draft, Washington, DC, 16 pp.
American Concrete Pavement Association (ACPA), 1991, “Design and
Tabatabie, A. M.; Barenberg, E. J.; and Smith, R. E., 1979, “Longi-
Construction of Joints for Concrete Highways,” ACPA, Skokie, IL, 24 pp.
tudinal Joint Systems in Slip-formed Rigid Pavements: V. II-Analysis
Albertson, M. D., 1992, “Fibercomposite and Steel Pavement Dowels,”
of Load Transfer Systems for Concrete Pavements,” Report No. DOT/
master’s thesis, Iowa State University, Ames, IA, 268 pp.
FAA RD-79/4, Federal Aviation Administration, U.S. Department of
ASTM D3171, 2011, “Standard Test Methods for Constituent Content of
Transportation, Washington, DC, 193 pp.
Composite Materials,” ASTM International, West Conshohocken, PA, 10 pp.
Westergaard, H. M., 1925, “Computation of Stresses in Concrete
ASTM D3418-12e1, 2012, “Standard Test Method for Transition
Roads,” Proceedings of the 5th Annual Meeting of the Highway Research
Temperatures and Enthalpies of Fusion and Crystallization of Polymers by
Board, Washington, DC, pp. 91-112.
Differential Scanning Calorimetry,” ASTM International, West Consho-
Westergaard, H. M., 1928, “Spacing of Dowels,” Proceedings of the
hocken, PA, 7 pp.
Highway Research Board, No. 8, Transportation Research Board, Wash-
ASTM D4475, 2008, “Standard Test Method for Apparent Horizontal
ington DC, pp. 154-158.
Shear Strength of Pultruded Reinforced Plastic Rods by the Short Beam
William, G. W., and Shoukry, S. N., 2001, “3D Finite Element Analysis
Method,” ASTM International, West Conshohocken, PA, 4 pp.
of Temperature Induced Stresses in Dowel Jointed Concrete Pavements,”
ASTM D4476, 2009, “Standard Test Method for Flexural Properties
International Journal of Geomechanics, V. 1, No. 3, pp. 291-307.
of Fiber Reinforced Pultruded Plastic Rods,” ASTM International, West
Yoder, E. J., and Witczak, M. W., 1975, Principles of Pavement Design,
Conshohocken, PA, 5 pp.
second edition, John Wiley & Sons, Inc., New York, 736 pp.
ASTM D4694, 2009, “Standard Test Method for Deflections with a
Falling-Weight-Type Impulse Load Device,” ASTM International, West
Conshohocken, PA, 3 pp.

ACI Structural Journal/March-April 2014 341


NOTES:

342 ACI Structural Journal/March-April 2014


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 111-S30

Nonlinear Static Analysis of Flat Slab Floors with Grid


Model
by Dario Coronelli and Guglielmo Corti
A grid model has been set up for the nonlinear response of a flat slab The test slab consists of a rectangular floor with 16 slab-
subjected to gravity and lateral cyclic loading. The model requires column joints under gravity and biaxial cyclic quasi-static
the definition of the grid geometry and properties of point hinges horizontal loading increased up to failure. Due to the
in beam finite elements, and modelling the nonlinear response different column cross sections and slab reinforcement, the
in bending, torsion, and shear. The simulation is carried out for
structure is symmetric in one direction and nonsymmetric in
experimental tests on a floor under gravity and lateral biaxial
the orthogonal direction.
cyclic loading of increasing amplitude. Pushover analyses have
been performed under gravity and horizontal loads in the two prin- The experimental response showed significant effects of
cipal directions. Predictions are shown of the global response and the biaxial loading.3 The maximum connection moments
the connections of different column shapes and slab reinforcement were reached at 4% drift in the north-south (N-S) direction,
with the strength, drift capacity, and failure modes. The accuracy is whereas in the east-west (E-W) direction, for most connec-
different in the two directions of loading due to the damage of the tions, the maximum moments were recorded at 2% drift and
test slab for biaxial cyclic loading. The results show the potential punching failures occurred for drifts at approximately 3%.
of the model for design and analysis of existing flat slab structures. Following this, the test was stopped. Biaxial loading effects
have also been documented in tests on connections,4 with
Keywords: flat slabs; grid models; punching shear; slab-column connections.
a reduction of drift capacity in both directions with respect
to the case of uniaxial loading. This effect has also been
INTRODUCTION
measured in connections with rectangular cross sections of
Reinforced concrete slabs, supported by columns and
the column.5,6
walls, are among the most common structures for floors, and
The first section of the paper briefly describes the exper-
their diffusion is continuously increasing all over the world.
imental tests.3 The second section presents the setup of the
Their use is advantageous for speed in construction, the
grid model; in particular, the definition of the grid geom-
possibility to realize flat intrados without beams, and their
etry, nonlinear hinges, and loads applied. In the third section,
low thickness in relation to the span.
results of nonlinear analyses under gravity and lateral loads
One of the leading aspects in slab design is to guarantee
are shown and compared with experimental results.3 The
adequate punching shear resistance of slab-column connec-
conclusions assess the adequacy of the modeling of the
tions when the structure is subjected to both gravity and
global response and of the individual connections, with
seismic loads. A wide variety of literature is available on
attention given to prediction of the ultimate load and drift
this topic, with code provisions accompanied by several
capacity, and the failure modes.
constructions in seismic zones.1
A grid model was previously developed and validated for
RESEARCH SIGNIFICANCE
the nonlinear behavior of flat slab-column connections.2 The
A grid model developed for slab-column connections is
grid is composed by linear beam finite elements; the inelastic
validated for the pushover analysis of a flat slab floor tested
response of the structure is concentrated in nonlinear point
experimentally under gravity and cyclic lateral loading. The
hinges. A model2 has been proposed for static pushover
performance for a nonsymmetrical geometry is studied,
analyses that permit evaluation of flexural, torsional, and
with four different types of column cross sections arranged
shear internal forces and moments in the whole plate struc-
nonsymmetrically with respect to one principal direction of
ture, with particular attention to the state of slab-column
the plan; the reinforcement layout varies accordingly. The
joints. The description of their nonlinear behavior allows
approximation of the slab behavior with biaxial load cycles
evaluation of the whole slab structural response up to failure.
of increasing amplitude up to failure is investigated, with a
In particular, it is possible to assess the safety of connec-
strong damage accumulation effect. The results corroborate
tions with respect to punching and structural deformability,
the potential of the model for design and analysis of existing
in terms of interstory drift ratio under lateral loads.
flat slab structures.
The aim of this paper is to show the efficiency of the model2
for flat slab floors under gravity and horizontal loads. The
grid model2 was used to simulate the experimental tests on
a scaled model of a flat-plate structure.3 A rather demanding ACI Structural Journal, V. 111, No. 2, March-April 2014.
MS No. S-2012-103.R2, doi:10.14359.516686526, was received April, 8 2013, and
case study, including structural irregularity and the effects of reviewed under Institute publication policies. Copyright © 2014, American Concrete
biaxial loading, was chosen. Institute. All rights reserved, including the making of copies unless permission is
obtained from the copyright proprietors. Pertinent discussion including author’s
closure, if any, will be published ten months from this journal’s date if the discussion
is received within four months of the paper’s print publication.

ACI Structural Journal/March-April 2014 343


Fig. 1—Geometry of test slab, plan view, in. (cm); posi-
tions of slab column connections (letters A through D and
numbers 1 through 4).3
EXPERIMENTAL RESULTS
The tests regard the scaled model of a flat-plate structure
subjected to gravity and lateral loads designed following
ACI 318-83.3 The prototype slab represents a portion of a
typical flat-plate floor of an intermediate story of a multi-
story office building. The slab has three bays in each direc-
tion, and a 203 mm (8 in.) thickness. The bay widths are Fig. 2—Experimental crack pattern of top and edge of slab
6.86 and 4.57 m (22.5 and 15 ft) for long and short direc- following test at 4% drift in E-W direction3 and represen-
tions, respectively. The story height is 3.05 m (10 ft). tation of the numerical deformations and position of the
The scaled model (Fig. 1) used for the experimental nonlinear hinges activated for W loading direction.
study has dimensions equal to 40% of those of the proto-
Testing of the model slab included gravity load tests,
type. The length of each bay is 2.74 and 1.83 m (108 and
construction load tests, and lateral load tests up to failure
72 in.) for the long and short directions, respectively. The
(Appendix B). A uniform gravity loading of 3.73 kN/m2
slab is 81 mm (3.2 in.) thick. Columns extend above and
(78 lb/ft2) was provided by weights placed on the slab,
below the slab. The column stubs above the slab were 0.3 m
with a total vertical uniform load with the self-weight of
(12 in.) long, and their purpose was to anchor the column
the structure equal to 5.65 kN/m2 (118 lb/ft2). Construction
longitudinal reinforcement and to provide continuity of the
loads were then applied in sequence on each panel of the
columns themselves through the floor; the inferior columns
model slab; their approximated value is equal to 2.63 kN/m2
stubs were 1.46 m (4.8 ft) long with pinned connections at
(55 lb/ft2), then removed for all subsequent tests. The total
the extremities.
vertical uniform load with the self-weight of 5.65 kN/m2
Four different column cross-sectional shapes were used
(118 lb/ft2) was used for the lateral load testing described in
to collect data related to the effects of column rectangu-
the following, resulting in a gravity shear ratio Vg/Vc close
larity on the structural response: rectangular columns with
to 0.2 (Table B2). The lateral loads were applied to the test
an aspect ratio of 2:1 in the east half of the floor and square
slab in two principal directions (N-S and E-W) using four
columns in the west half. With this layout, the structure is
reversible hydraulic actuators supported on a reaction frame.
symmetric about the floor centerline along the long direction
Increasing lateral drifts from 1/400 to 1/25 were imposed,
and nonsymmetric in the orthogonal direction.
with lateral drift first imposed in the S direction, then
The concrete had a mean cylinder compressive strength
reversed to the N direction, and finally returned to zero drift.
of 21.8 MPa (3160 psi). The self-weight of the slab was
The cycle was repeated once. The same sequence was then
1.9 kN/m2 (40 lb/ft2), corresponding to 23.6 kN/m3
immediately applied in E-W direction. These alternating
(150 lb/ft3). Slab longitudinal reinforcement (Appendix A*)
loadings in the two orthogonal directions were then repeated,
was made of No. 2 deformed bars with cross-sectional area
doubling the lateral drift (Fig. B1). No torsional effects
Al = 32 mm2 (0.05 in.2) (Φ = 6.4 mm [0.25 in.]), with fy =
leading to rotation of the floor were observed in the tests.
462.6 MPa (67.1 ksi). The layout of the reinforcement reflects
The experimental response to the biaxial horizontal loading
the changes in the cross sections of the columns (Appendix A).
showed relevant differences in the two orthogonal directions
No transverse reinforcement was used for the slab.
as a consequence of the loading history.3 The connections
*
in the N-S direction of loading reached their strength at 4%
Appendixes are available at www.concrete.org/publications in PDF format,
appended to the online version of the published paper. It is also available in hard copy drift. Loading in the E-W direction then took place, and
from ACI headquarters for a fee equal to the cost of reproduction plus handling at the the expected connection strengths were not developed. The
time of the request.

344 ACI Structural Journal/March-April 2014


recorded maximum moments for most connections were
those of the previous load cycle at 2% drift, followed by
severe strength deterioration and punching failures around
3% drift. The crack pattern of the entire slab at the end of
these tests is shown in Fig. 2.

NUMERICAL MODEL
In the following, the description of the setup of the grid
model for the particular study case previously explained and
the comparison between test results and numerical outputs
will be presented. The grid model2 was developed using a
commercial finite element software7 to perform nonlinear
static analyses under displacement control. The slab is repre-
sented by a grid of linear beam finite elements, fixed at joints,
arranged in two orthogonal directions; the columns are
modeled with two beam elements, one above and one below
the slab level, fixed to the plate. The beam finite elements
have been defined as beam-column elements, thus including
the effects of flexural, torsional, and shear deformations.
Each joint has six degrees of freedom. The in-plane shear
due to the lateral loads is modeled by the flexural stiffness of
the elements around the axis perpendicular to the slab plane,
together with the in-plane shear and axial stiffness.
Grid geometry—To design the grid geometry, guidelines
given by CIRIA Report 1108 have been considered as a
reference point, resulting in the choices for the grid detailed
in the following. The grid spacing must be sufficiently close
near to the columns to obtain a good approximation of the
load effects in the slab, since concentration of internal forces
and moments exists in these zones; the elements can be more
widely spaced elsewhere. For the size of the grid spacing,
generally a width equal to c + d is adequate,2 where c is the
column dimension, and d is the effective height of the slab.
It should be noted that c + d is the width of the shear crit-
ical section of connection, as considered in the grid model2
according to the definition of ACI 3189; further details are Fig. 3—Grid model geometry: (a) overall view of grid and
given in the following section. columns; and (b) detail of model at connection of slab
Observing the layout of the test slab (Fig. 1), it is evident and column.
that the structure is asymmetrical in regards to the column
Columns sections are taken with the same dimensions of
sections; four different column sections were used by
those effectively used in the test slab.
authors of the experimental test.3 For beam elements placed
Loads—The structure is subjected to gravity and lateral
on the axes connecting the connections, a width equal
loads. From the total uniformly distributed gravity load per
to the major c + d along the axes themselves was chosen
unit of surface (comprising self-weight), the total load per
(Appendix B). The nearby elements have the same width,
unit length acting in the two principal directions is calculated
with some adjustments to have the sum of the widths of all
for each element of the grid. With regard to the imposed
cross sections equal to the slab dimensions in the plan. The
displacements, these have been applied in the same positions
depth of the beam elements modeling the slab is equal to the
of the experimental test compatibly to the refinement of the
slab thickness.
grid. A displacement control analysis is carried out.
The portions of grid corresponding to the column cross-
Nonlinear hinges—Each grid element is composed of
section dimensions are modeled by four beam elements
an elastic part and nonlinear hinges. The cracked stiffness,
positioned along the four column semi-axes. A high stiff-
depending on the quantity of reinforcement, is used for the
ness is specified to these elements, to model the support given
elastic part of all elements to consider the effect of shrinkage
to the slab by the column. In the analyses, the elements had
and construction stresses,3 causing cracking in the slab
the column width and a depth equal to three times the slab
previous to the action of lateral loads. The hinges are points
thickness.
in which the nonlinear properties of the elements them-
As already specified, columns are modeled with two beam
selves are lumped, and are characterized by relations that
elements: one below the slab grid with a length of 1.260 m
link bending moment, shear, and torsion with the inelastic
(4.13 ft), and one above with a length of 0.345 m (1.13 ft).
curvature, shear distortion, and twist angle, respectively.
The columns stubs below the slab are pinned at their bases.

ACI Structural Journal/March-April 2014 345


the optimal values for modeling the experimental punching
in tests carried out on slab-column connections,2 with γu =
0.0092[pfy/(1/3fc0.5)] – 0.011, where p is the longitudinal
reinforcement ratio and fc′ and fy are the strength of mate-
rials. The expression in customary units is given by γu =
0.00229[pfy /(fc′0.5)] – 0.011, where fc′ and fy are expressed
in psi.
A trilinear relation was defined for the shear force-strain
relation (Fig. 4); the behavior is obtained with the sectional
shear model,10 normalized with respect to the maximum
shear force and the correspondent strain. The normalized
diagrams with the associated shear capacity Vu and the corre-
sponding ultimate shear distortion γu are used as input to the
Fig. 4—Shear nonlinear hinge: relation of shear force and model. The first point of the trilinear diagram corresponds
inelastic strain. to cracking. The second point (γ1, V1/Vu) is determined by
the intersection of the portion in the cracked stage of the
The elements (Fig. 3(b)) are made of a linear elastic beam curve determined by the sectional model10 with the level of
element with a nonlinear hinge at each end for flexure. Each 80% of the strength. The third point corresponds to the shear
element has one hinge for shear and one for torsion, both strength, followed by the last point at the end of the soft-
at the center of the element. The elements are connected ening branch determined using the shear sectional model.10
together at their end nodes in the intersection points of the The third point corresponds to the shear strength, followed
grid. The simplified phenomenological approach adopted by the last point at the end of the softening branch deter-
to define the hinge properties is summarized herein; more mined using the shear sectional model.10 The hinge length is
details can be found in the references.2 taken equal to 2d.2 Shear nonlinear hinges were applied only
Point hinges were assigned to frame elements lying on to the grid elements framing into the columns; this choice is
the center lines connecting joints and on the immediately based on the fact that shear forces in other elements of the
adjacent axes. This choice was made to make the model grid are much lower.
more computationally light. The level of internal forces was Torsion—The definition of torsion nonlinear hinges in the
checked a posteriori, indicating that the cracking moments grid model2 relates torque and twist; normalized diagrams
were not reached in the internal part of the slab panels. with respect to the ultimate torsional capacity Mtu and the
Three types of hinges are defined according to the grid corresponding ultimate twist ψu are provided. Mtu0 is the
model2: flexural, shear, and torsional. Their properties torsional capacity without interaction effects, and Mtu
depend on the geometry of each element’s cross section, on considers the interaction of the torsional strength with the
concrete properties, and on longitudinal bottom and top rein- bending moment and shear acting in the connection due
forcement of the slab. to gravity loads. These are determined according to the
Bending—Flexural nonlinear hinge properties are model for slab-column connections proposed by Park and
computed analytically for each beam element using a Choi,12 where Mtu0 is the product of the polar moment of
sectional model10 with nonlinear constitutive relations inertia at the sides [J′ = 2(c + d)d3/12 + 2d(c + d)3/12] by
based on perfect bond and plane section assumptions; the a strength vus = 5.0vn. To consider the interaction with the
moment-curvature relations thus obtained are then approx- bending moment and shear acting in the connection due to
imated with trilinear relationships. Normalized diagrams gravity loads, a reduced torsional capacity Mtu is calculated
with respect to ultimate capacity Mu and associated ultimate by reducing the strength to vus,red = vn[5.0 – 2.5(σn/fc)] –
curvature Φu are provided as input to the model, together Vg/(c + d)/d, where σn is the maximum flexural stress in the
with the values of Mu and Φu. The hinge length is defined concrete at the sides of the connection and Vg/(c + d)/d is the
equal to d, based on test results.2 gravity shear stress. With regard to the ultimate twist angle
Shear—The shear force-distortion relation is input as ψu0, values measured experimentally for specimens without
a normalized diagram with respect to the ultimate shear transverse steel13 are used2; a value of 0.02 is adequate.14
capacity Vu and to the associated ultimate distortion γu When the interaction with shear and bending moment deter-
(Fig. 4). The shear capacity of each side of the critical mines a capacity reduction from Mtu0 to Mtu, the twist angle
perimeter was computed as Vu = vn(c + d)d according to ψu0 is reduced to ψu, proportionally to Mtu/Mtu0.
ACI 318-05,9 with the strength vn = 1/3(fc)1/2 MPa [4(fc′)1/2 For the normalized torque and twist relation, a trilinear
psi] for elements without shear reinforcement. Experiments diagram is used; the behavior is obtained with the torsional
show that this can vary with the slab thickness, reinforce- model of Collins and Mitchell,15 normalized with respect to
ment ratio, and the effects of cyclic loading.11 For this the maximum torque and the correspondent twist. The piece-
case, the slab had a low thickness that would increase vn; wise linear diagram is obtained by connecting the points at
the cyclic loading could cause strength deterioration. Hence cracking, yielding, and ultimate (Fig. 5).
vn = 1/3(fc)1/2 [4(fc′)1/2 psi] was used for the analyses. The Torsional nonlinear hinges computed by the afore-
values of ultimate shear distortion γu for slabs without shear mentioned method are used in elements framing into the
reinforcement were obtained by a linear interpolation of columns. For grid elements that are not at the interface with

346 ACI Structural Journal/March-April 2014


Fig. 5—Torsion nonlinear hinge: relation of torque and
inelastic twist angle.

Fig. 7—Comparison between numerical pushover curve


and experimental lateral load-drift envelope curve for E-W
direction (experimental cycles in small figure)
directions of loading. The experimental responses in the
two orthogonal directions are shown in Fig. 6 and 7. As a
consequence of the orientation of a part of the columns with
the strong axis in the E-W direction, the sum of the column
stiffness in the E-W direction is the double that in the N-S
direction. This has a limited impact on the global stiffness
of the structure resulting from the assemblage of columns
and slab. For the first cycles at low drift, the stiffness in the
E-W direction was only 20 to 30% higher than in the N-S
direction (Table B1). The curves up to 1% drift are rather
close in the two directions, whereas they differ both in terms
of resistance and ultimate drift. As previously mentioned,
Fig. 6—Comparison between numerical pushover curve and this difference can be explained by the degradation of slab-
experimental lateral load-drift envelope curve for N-S direc- column connections due to the biaxial cyclic loading.3,4
tion (experimental cycles in small figure). In Fig. 6 and 7, the model predicts an ultimate drift capacity
of 4% in both directions. The prediction in the N-S direction
the columns, the model prescribes an elastic-perfectly plastic is accurate; a slight difference shows in the N direction with
behavior, with a torsional strength16 equal to 0.58(fc′)0.5 MPa negative forces. It should be considered that experimen-
[6.99(fc′)0.5 psi]. tally, the – 4% drift was reached after that the same drift was
attained in the S direction, with some damage accumulation.
Comparison between numerical and experimental In the E-W directions, differences between tests and model
results response show starting from 2% drift. The ultimate drift and
The procedure followed for the analysis of the model slab load capacity are overestimated, with the model reaching
is the following: a load-controlled nonlinear analysis is first ultimate drift at 4%; the tests showed strength deteriora-
performed under gravity loads specified previously; then, tion beyond 2% drift, followed by several punching failures
starting with these results, nonlinear static analyses for hori- at drifts around 3% on average. The high level of damage
zontal loads with displacement control up to failure in each inflicted by the two cycles at 4% drift in the N-S direction
principal direction are carried out. It should be noted that should be considered. This is not taken into account by
both positive (S and W) and negative (N and E) verses of the model. A similar analytical overestimation of the load
loading are considered in the analyses, with a total of four capacity was obtained by the authors of the tests3 using the
analyses for horizontal load, each starting after the gravity ACI 318-83 model for the strength of the connections; the
load analysis. experimental interior connection strengths were, on average,
The experimental study3 provides data for the total lateral 87% of the analytical values in the E-W direction, whereas
load and drift applied to the structure (Fig. 6 and 7). Enve- the model was slightly conservative (experimental results
lopes of peak values of the lateral loads and the corresponding are 110% on average of the analytical) in the N-S direction.
deflections in each cycle are well suited to be compared The model approximates the response, taking into account
with pushover curves of the model for the two orthogonal the global behavior of the structure with different column

ACI Structural Journal/March-April 2014 347


Fig. 8—Comparison between numerical and experimental moment-rotation curve of Connection C3 for: (a) N-S; and (b) E-W
directions. Positive drifts S and W. Experimental envelope up to connection strength. Experimental E-W 4% drift cycles,
including cycles beyond failure.
cross-section shapes, with the nonsymmetry highlighted in With regard to the interior Connection C3 and considering
the description of the structure. the N-S loading direction (Fig. 8(a)), the numerical curve
Figure 2 shows the experimental crack pattern at failure fits very well the experimental one for positive and negative
with the slab numerical deformed shape and the nonlinear verse of loading. For the E-W loading direction (Fig. 8(b)),
hinges activated in the elements for the W loading direction. a fair approximation is obtained, though some difference
The most damaged zones—that is, where hinges are beyond between the numerical and the experimental curve shows
yielding—are those around connections corresponding to from approximately 1.5 to 2% drift, overestimating the
the experimental damage. A more detailed analysis of the strength; this is due to the previously mentioned deteriora-
prediction of damage is carried out as follows. tion of the test slab subjected to the biaxial cyclic loading.
Punching failure of a connection is reached after the The model reproduces the connection failure modes that
formation of a shear hinge, when the torsion capacity at were experimentally observed. Connection C3 failed for
the sides of the connection is exceeded. In the analyses punching shear during the final test in the E-W direction,
presented herein, the structure reached failure with punching for both positive and negative directions of loading at a drift
in part of the connections; several others were very close of 3.1%. Numerical results effectively confirm these data,
to this stage. This reflects the experimental behavior where because punching of the connection is not detected during
failure occurred with punching occurring in several connec- analyses for the N-S loading direction (Fig. 9, left), whereas
tions within a short interval of time.3 punching is predicted in the E-W loading direction (Fig. 9,
right) for both positive and negative drifts of 3.9 and 4.0%,
Unbalanced moment versus rotation envelopes respectively. The difference between experimental and
of connections numerical punching drift is ascribed to the biaxial loading.
The behaviors of each connection obtained numeri- For the lateral Connection D3, when loaded parallel to the
cally and experimentally3 were compared. The test report3 free edge (N-S), good correspondence is obtained with the
provides unbalanced moment-rotation envelopes for each pushover (Fig. 10(a)) in terms of force and drift. For the E-W
joint up to the drift corresponding to connection strength cycles, the experimental failure was punching in both E and
for both the loading directions. Experimental cycles for the W directions at 3 and 2% drift, respectively. Strength is over-
E-W 4% drift test are also presented, including the cycles estimated in the W loading (Fig. 10(b)) at a maximum drift
beyond failure. For the model, the moments were obtained very close to the experimental. Punching failure is predicted
by the product of the numerical base shear and the column in the E (negative) direction (Fig. 11, right), though strength
height3; rotations are an output of the analysis in the nodes and drift capacity are overestimated.
connecting the slab and columns. The behavior of connec- With regard to corner connections, for A4 (Fig. 12), the
tions depends on their position (internal, edge, and corner) numerical curve predicts drift capacity accurately in the N-S
and the direction of loading (N-S and E-W). and W directions, but overestimates the drift capacity in the
The figures in the following show examples of the compar- E (negative) direction. The connection strength is overesti-
ison of numerical and experimental moment-rotation curves mated in the N (negative) and E-W directions.
of connections, representative of three different typology The tests showed that all corner connections, including
of joints, discussed previously: the interior Connection C3 A4, survived the E-W 4% drift test, except D1 in the W direc-
(Fig. 8 and 9), the lateral Connection D3 loaded parallel to tion. The numerical model correctly shows that punching is
the free edge (Fig. 10 and 11), and the corner Connection A4 not reached, whereas the torsion capacity is reached for both
(Fig. 12 and 13). S and W directions, corresponding to the experimental crack
patterns (Fig. 13).

348 ACI Structural Journal/March-April 2014


Fig. 9—State of hinges around column of Connection C3 for N-S (left) and E-W (right) loading at ultimate drift. Experimental
crack pattern on top surface after punching (with line surrounding damaged zone).

Fig. 10—Comparison between numerical and experimental moment-rotation envelope of Connection D3 for: (a) N-S; and
(b) E-W direction. Positive drifts S and W. Experimental envelope up to connection strength. Experimental E-W 4% drift cycles,
including cycles beyond failure.
Summing up, the model provides a good approximation accumulation of damage with biaxial loading, and the model
of the experimental behavior for the N-S direction, where provides non-conservative results. The same results were
the biaxial loading effects were lower. In the E-W direction, reached for strength of the connections using ACI 318-83,
the strength and ultimate drift capacity were affected by the

ACI Structural Journal/March-April 2014 349


and showing the experimental ultimate drift capacities
reduction in the E-W direction compared with N-S.3

CONCLUSIONS
A grid model has been set up to reproduce the nonlinear
response of a flat slab structure subjected to gravity and lateral
cyclic loading. Experimental tests carried out on a scaled
model were analyzed. The slab was tested under gravity and
lateral biaxial cyclic loading. Both principal directions were
loaded alternatively, with a sequence of cycles of increasing
amplitude. Nonlinear static analyses under gravity loads
have been performed, followed by pushover analyses under
horizontal loads in the two principal directions of the slab on
the numerical model.
The model is efficient in showing several aspects of the
response of the test slab considered:
1. The experimental global behavior is approximated
differently in the two orthogonal directions of biaxial
loading: the pushover curves in N-S directions are very close
to the experimental in terms of path, maximum load, and
maximum drift; for the E-W direction, a numerical overes-
timation of lateral load and drift is detected that is due to
the experimental degradation of the test slab due to biaxial
cyclic loading.
2. The model captures the ultimate drift capacity of the
test slab in both N and S directions, the first loaded up to 4%
drift. The drift capacity in the E and W direction is overes-
timated by the model, predicting a maximum drift close to
4% in both E and W directions against experimental values
close to 2%.
3. Experimental moment-rotation curves of internal
connection are well approximated by numerical curves for
the N-S loading direction; for the E-W direction, an overesti-
mation of load and drift capacity is detected. A similar result
is obtained for lateral connections loaded parallel to the free
edge. The difference is attributed to the slab degradation
Fig. 11—State of hinges around column of Connection D3 observed experimentally, due to the damage accumulation
for N-S (left) and E-W (right) loading at ultimate drift. with biaxial loading.
4. For lateral connections loaded perpendicular to the
free edge and corner connections, the numerical analysis is
less accurate for both the N-S and E-W directions. In the

Fig. 12—Comparison between the numerical and experimental moment-rotation envelope of joints A4 for: (a) N-S, and
(b) E-W direction. Positive drifts S and W. Experimental envelope up to connection strength. Experimental E-W 4% drift
cycles, including cycles beyond failure.

350 ACI Structural Journal/March-April 2014


Guglielmo Corti is a Civil Engineer. He received his MSc in civil engi-
neering from Politecnico di Milano in 2010. His research interests include
design of reinforced concrete structures.

NOTATION
c = column side
d = slab average effective depth
J′ = polar moment of inertia at sides of connection
Mtu = reduced torsion capacity (with interaction effects)
Mtuo = torsion capacity without interaction effects
Mu = flexural capacity
Vc = punching shear capacity of flat slab-column connection
Vg = shear force transferred between slab and column under gravity
loads
Vu = shear force at capacity of critical sections
vc = concrete shear strength, ACI 318-059
vn = shear strength, ACI 318-059
vus = eccentric shear strength at the sides of connection
vus,red = reduced eccentric shear strength, by interaction with flexure and
shear
Φu = ultimate curvature
σn = maximum flexural stress in concrete, compression
γu = inelastic shear strain corresponding to shear capacity
ψu = reduced twist angle at maximum torque (with interaction effects)
ψuo = twist angle at maximum torque

REFERENCES
1. Joint ACI-ASCE Committee 421, “Seismic Design of Punching Shear
Reinforcement in Flat Plates (ACI 421.2R-07),” American Concrete Insti-
tute, Farmington Hills, MI, 2007, 24 pp.
2. Coronelli, D., “A Grid Model for Flat Slab Structures,” ACI Structural
Journal, V. 107, No. 6, Nov.-Dec. 2010, pp. 645-665.
3. Hwang, S. J., and Moehle, J. P., “An Experimental Study of Flat-
Plate Structures under Vertical and Lateral Loads,” Technical Report UCB/
EERC-93/03, Earthquake Engineering Research Center, University of Cali-
Fig. 13—State of hinges around column of Connection A4 fornia at Berkeley, Berkeley, CA, Feb. 1993, 278 pp.
4. Pan, A. D., and Moehle, J. P., “Lateral Displacement Ductility of
for N-S (left) and E-W (right) loading at ultimate drift. Reinforced Concrete Flat Plates,” ACI Structural Journal, V. 86, No. 3,
Experimental crack patterns (side and top). May-June 1989, pp. 250-258.
5. Tan, Y., and Teng, S., Interior Slab-Rectangular Column Connec-
latter case, the accumulation of damage with the biaxial tions under Biaxial Lateral Loadings, SP-232, M. A. Polak, ed., American
Concrete Institute, Farmington Hills, MI, 2005, pp. 147-174.
loading affects the results. Improvements are needed for the 6. Anggadjaja, E., and Teng, S., “Edge-Column Slab Connections
parameters in the model for the strength of these types of under Gravity and Lateral loading,” ACI Structural Journal, V. 105, No. 5,
connections. Sept.-Oct. 2008, pp. 541-551.
7. CSI, CSI Analysis Reference Manual—SAP 2000 Advanced Research
5. The model reproduces the experimental failure modes v.10, Computers and Structures Inc., Oct. 2005, 433 pp.
of connections. The tests considered herein showed the 8. Whittle, R. T., “Design of Reinforced Concrete Flat Slabs to BS
effects of biaxial loading as a reduction of ultimate drift 8110,” Report 110, CIRIA, 1994.
9. ACI Committee 318, “Building Code Requirements for Structural
and capacity at punching of the slab-column connections, Concrete (ACI 318-05) and Commentary (318R-05),” American Concrete
with respect to those for a one-directional loading. These Institute, Farmington Hills, MI, 2005, 430 pp.
punching failures are predicted by the model in correspon- 10. Bentz, E. C., Sectional Analysis of RC Members, University of
Toronto, Toronto, ON, Canada, 2000, 198 pp.
dence of ultimate moment and drift values higher than the 11. Dilger, W. H., “Flat Slab-Column Connections,” Progress in Struc-
experimental. The developments of the research should tural Engineering and Materials, V. 2, 2000, pp. 386-399.
improve the model to consider the biaxial loading effect. 12. Park, H., and Choi, K., “Improved Strength Model for Interior Flat
Plate-Column Connections Subjected to Unbalanced Moment,” Journal of
These results show the performance of the model in a Structural Engineering, ASCE, V. 132, No. 5, May 2006, pp. 694-704.
rather demanding case study, including some structural 13. Kanoh, Y., and Yoshizaki, S., “Strength of Slab-Column Connec-
irregularity and the effects of cyclic loading in two orthog- tions Transferring Shear and Moment,” ACI Journal, V. 76, No. 3, Mar.
1979, pp. 461-468.
onal directions. The results for biaxial loading indicate the 14. Tian, Y.; Jirsa, J. O.; and Bayrak, O., “Nonlinear Modeling of Slab-
need for further research developments. Column Connections under Cyclic Loading,” ACI Structural Journal,
V. 106, No. 1, Jan.-Feb. 2009, pp. 30-38.
15. Collins, M. P., and Mitchell, D., “Shear and Torsion Design of
AUTHOR BIOS Prestresses and Non-Prestressed Concrete Beams,” PCI Journal, Sept.-Oct.
Dario Coronelli is an Assistant Professor in the Department of Civil and 1980, pp. 32-100.
Environmental Engineering at Politecnico di Milano, Milan, Italy, where 16. Park, R., and Paulay, T., Reinforced Concrete Structures, John
he received his PhD in 1998. His research interests include the structural Wiley & Sons, Inc., New York, 1975, 769 pp.
effects of corrosion and seismic design of reinforced structures.

ACI Structural Journal/March-April 2014 351


NOTES:

352 ACI Structural Journal/March-April 2014


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 111-S31

Effect of Steel Stirrups on Shear Resistance Gain Due


to Externally Bonded Fiber-Reinforced Polymer Strips
and Sheets
by Amir Mofidi and Omar Chaallal
This paper presents the results of an experimental investiga- on the contribution of FRP to shear resistance. The effect of
tion of reinforced concrete (RC) T-beams strengthened in shear internal transverse steel reinforcement is one of these influ-
with externally bonded (EB) carbon fiber-reinforced polymer ential parameters. The adverse effect of transverse steel on
(CFRP) strips and sheets. The main objective of this study was to the effectiveness of externally bonded FRP used for shear
gain insight, by varying the test parameters, into the interaction
strengthening and retrofit of RC beams is well established
between the internal transverse steel reinforcement and the exter-
(Bousselham and Chaallal 2008; Chaallal et al. 2011; Mofidi
nally bonded CFRP used for strengthening of RC beams in shear.
The test parameters of this study were: 1) the CFRP ratio (that and Chaallal 2011a,b). In contrast, the effect of adding EB
is, the spacing of the CFRP strips); 2) the presence or absence of FRP for shear retrofit on the performance of internal steel
transverse steel; 3) the transverse steel ratio (that is, the spacing stirrups has not been thoroughly documented. Moreover,
of the stirrups); and 4) the use of CFRP strips versus CFRP sheets. in most modern codes and guidelines, the contribution of
In total, 10 tests were performed on 4520 mm (14 ft, 10 in.) long steel stirrups to the shear resistance Vs at the ultimate state is
T-beams. The study showed that the presence of internal transverse calculated on the premise that the steel stirrups have yielded.
steel reinforcement resulted in a significant decrease in the gain The premature debonding failure observed in RC beams
due to CFRP for all strengthened specimens. It also showed that strengthened in shear with FRP, however, has prompted
the steel yielded before failure for all test specimens with trans- legitimate questions and concerns as to whether the assump-
verse steel. Finally, the presence of CFRP did not result in a
tion that the steel stirrups yield before failure holds true
significant decrease in transverse steel strain. It can be concluded
(Chen et al. 2010).
that the contribution of transverse steel to shear resistance is not
affected by the addition of EB CFRP. These results are in good
agreement with the assumptions made by existing codes and design RESEARCH SIGNIFICANCE
guidelines, which are based on the yielding of transverse steel at Researchers have not yet reached a commonly accepted
ultimate strain for RC beams strengthened in shear with EB CFRP. agreement on the effect of transverse steel in RC beams retro-
fitted with FRP. Therefore, this effect is not yet considered
Keywords: carbon fiber-reinforced polymer (CFRP); CFRP sheet; in the design codes and guidelines, including ACI 440.2R-08
reinforced concrete beam; shear strengthening; strip; transverse
(ACI Committee 440 2008). Consequently, current design
steel reinforcement.
codes and guidelines may overestimate the shear resistance
of RC beams with transverse steel that are strengthened
INTRODUCTION
using EB FRP sheets and strips. Such uncertainties in shear
In recent years, shear strengthening of reinforced concrete
strengthening of RC beams using EB FRP have provided the
(RC) beams with externally bonded (EB) fiber-reinforced
key impetus for conducting the current research study, the
polymer (FRP) material has attracted attention and has been
objective of which was to gain insight into the interaction
studied by several researchers (Uji 1992; Chaallal et al.
between internal transverse steel reinforcement and exter-
1998; Triantafillou 1998; Khalifa et al. 1998; Bousselham
nally bonded FRP strips and sheets used for shear strength-
and Chaallal 2004; Chaallal et al. 2011; Mofidi and Chaallal
ening of RC beams. This insight has been achieved based on
2011a,b). These experimental and analytical studies have
results obtained from an experimental program carried out
provided valuable insights and results. Several questions,
on full-size T-beam specimens, described as follows.
however, still linger in the area of shear strengthening of
RC beams with FRP composites (Bousselham and Chaallal
EXPERIMENTAL PROGRAM
2004; Mofidi and Chaallal 2011a).
The experimental program (Table 1) involved 10 tests
For instance, a comparison between the experimental shear
performed on full-scale RC T-beams. The control spec-
resistance due to FRP and the values predicted by the analyt-
imens, which were not strengthened with carbon FRP
ical models used in existing codes and guidelines shows that
(CFRP), were labeled NF (for No FRP), whereas the speci-
major aspects of shear strengthening with FRP material are
mens retrofitted with a layer of EB CFRP sheet were labeled
still not captured by the predictive models used in the codes
and guidelines (Bousselham and Chaallal 2008; Mofidi and
ACI Structural Journal, V. 111, No. 2, March-April 2014.
Chaallal 2011a). This is mainly because the calculated shear MS No. S-2012-104.R2, doi:10.14359.51686527, was received November 12,
2012, and reviewed under Institute publication policies. Copyright © 2014, American
contribution of FRP according to the codes and guidelines Concrete Institute. All rights reserved, including the making of copies unless
does not account for the effect of certain parameters that permission is obtained from the copyright proprietors. Pertinent discussion including
author’s closure, if any, will be published ten months from this journal’s date if the
have experimentally been found to have a major influence discussion is received within four months of the paper’s print publication.

ACI Structural Journal/March-April 2014 353


Table 1—Experimental results
Resistance
Load at Total shear due to Deflection at
rupture, resistance, concrete, Resistance due Resistance due Gain due to loading point,
Specimen FRP type wf /sf kN kN kN to steel, kN to CFRP, kN CFRP, % mm
NR-NF — 0 122.7 81.2 81.2 0.0 0.0 0 2.6
NR-ST-LF Strips 87.5/175 203.1 134.5 81.2 0.0 53.3 66 6.2
NR-ST-HF Strips 87.5/125 227.3 150.6 81.2 0.0 69.3 85 7.2
NR-SH Sheet 1 181.2 120.0 81.2 0.0 38.7 48 4.2
HR-NF — 0 350.6 232.2 81.2 151.0 0.0 0 11.9
HR-ST-LF Strips 87.5/175 372.5 246.7 81.2 151.0 14.5 6 15.9
HR-ST-HF Strips 87.5/125 383.4 253.9 81.2 151.0 21.7 9 15.7
HR-SH Sheet 1 378.3 250.6 81.2 151.0 18.4 8 15.2
MR-NF — 0 294.0 194.7 96.2 98.5 0 0 11.2
MR-SH Sheet 1 335.2 222.0 96.2 98.5 27.3 14 11.3

Notes: 1 mm = 0.0394 in.; 1 kN = 0.225 kip.

Fig. 1—Test setup configuration: (a) cross section of tested RC beams; and (b) side view of loading configuration. (Note: 1 mm =
0.0394 in.)
SH (for SHeet), and the specimens strengthened with FRP 25.2 mm [1 in.], area 500 mm2 [0.78 in.2]) laid in two layers
strips (strip width = 87.5 mm = 3-7/16 in.) were labeled ST at the bottom, and six 10M bars (diameter 10.3 mm [ 0.4 in.],
(for STrips). Specimens strengthened with narrowly spaced area 100 mm2 [0.16 in.2]) laid in one layer at the top. The
FRP strips (spacing equal to 125 mm [5 in.]) were labeled bottom bars were anchored at the support with 90-degree
HF (for Heavily strengthened with FRP), whereas the spec- hooks to prevent premature anchorage failure. The internal
imens strengthened with widely spaced strips (spacing steel stirrups (where applicable) were 8 mm (5/16 in.) in
equal to 175 mm [6 7/8 in.]) were labeled LF (for Lightly diameter (area 50 mm2 [0.08 in.2]).
strengthened with FRP). Series NR (Not Reinforced with To apply the EB FRP sheets and strips to the RC spec-
transverse steel) consisted of specimens with no internal imens, the following steps were implemented: 1) the area
transverse steel reinforcement (that is, no stirrups). Series of the specimens where the CFRP sheets and strips was to
HR (Heavily Reinforced with transverse steel) and MR be epoxy-bonded was sand-blasted to remove any surface
(Moderately Reinforced with transverse steel) contained cement paste and to round off the beam edges; 2) the spec-
specimens with internal transverse steel stirrups spaced imen corners were chamfered to provide a radius of 12.7 mm
at s = d/2 and s = 3d/4, respectively, where d = 350 mm (0.5 in.) to avoid stress concentration in the FRP sheets
(13-3/4 in.) represents the effective depth of the cross section during the tests; 3) residues were removed using compressed
of the beam. Therefore, for instance, Specimen NR-ST-HF air; and 4) layers of U-shaped CFRP sheets and strips were
featured a beam with no transverse steel retrofitted using glued to the bottom and lateral faces of the RC beam using a
CFRP strips spaced at 125 mm (5 in.). The specimen details two-component epoxy resin.
are provided in Table 1, together with the identification
codes used hereafter. Materials
A commercially available concrete delivered to the
Description of specimens structural laboratory by a local supplier was used in this
The T-beams were 4520 mm (14 ft, 10 in.) long, and their project. The average 28-day concrete compressive strength
T-sections had overall dimensions of 508 x 406 mm (20 x was 25 MPa (3626 psi), which is very close to the average
16 in.). The width of the web and the thickness of the flange compressive strength of 27 MPa (3916 psi) obtained
were 152 and 102 mm (6 and 4 in.), respectively (Fig. 1(a) during the tests. It should be noted that the specimens of
and (b)). It should be noted that the web of the strengthened the MR series were cast using a different concrete batch,
beams is chamfered at the outer corners. The longitudinal the compressive strength of which was 35 MPa (5076 psi).
steel reinforcement consisted of four 25M bars (diameter

354 ACI Structural Journal/March-April 2014


location as the transverse steel. The load was applied using
a 2000 kN (449 kip) capacity MTS hydraulic jack. All tests
were performed under displacement-control conditions at a
rate of 2 mm/min (approximately 3/16 in./min).

ANALYSIS OF RESULTS
All the specimens failed in shear. The control specimens
failed due to diagonal tension failure of the concrete cross
section. The specimens strengthened with CFRP failed by
premature FRP debonding followed by diagonal tension
failure (Fig. 4(a) through (d)). Local CFRP fracture was
observed in few specimens (NR-ST-LF and NR-ST-HF);
this local failure is attributed to stress concentration at the
web corners.

Fig. 2—Side view of strengthened specimen using FRP sheet Deflection response
U-jacket with crack gauges on CFRP strips. Figure 5 compares the deflection response for RC beams
without transverse steel reinforcement. It reveals that the
The scatter between the results of compression tests on the NR-SH and NR-ST-HF specimens exhibited slightly greater
cylinder specimens was insignificant. overall stiffness than the other beams. Specimen NR-ST-HF
The longitudinal steel reinforcement consisted of exhibited the highest deflection at the loading point and a
25M bars (modulus of elasticity 187 GPa [27,122,057 psi], higher maximum load at failure than the other specimens
and yield stress 500 MPa [72,519 psi]), and the transverse (Fig. 5). The beams strengthened with FRP strips exhibited
steel reinforcement consisted of deformed 8 mm (13/16 in.) more deformability than the beams strengthened with FRP
bars (modulus of elasticity 206 GPa [29,877,774 psi] and sheets. This occurred mainly because in RC beams strength-
yield stress 650 MPa [94,275 psi]). ened with FRP strips, local FRP debonding did not result
The composite material was a unidirectional carbon-fiber in a complete debonding failure. Each local strip-debonding
fabric epoxy-bonded over the test zone in a U-shape around event resulted in a drop in the load-carrying capacity of the
the web (Fig. 2). The dry CFRP sheet had an ultimate tensile beam (Fig. 5), but the load continued to increase as the cracks
strength of 3450 MPa (500,380 psi), an elastic modulus of propagated in the RC beams web, engaging the unloaded
230 GPa (33,358,697 psi), and an ultimate strain of 1.5%, CFRP strips in their path. In specimens strengthened with
as reported by the manufacturer. The thickness of the CFRP FRP strips, and unlike beams strengthened with FRP sheets,
fabric used was 0.11 mm. local debonding of FRP cannot propagate from one FRP
strip to the next. Therefore, using FRP strips results in a
Test setup more progressive type of failure, and a sudden and brittle
All 10 tests were conducted in three-point load flexure failure is prevented.
(Fig. 1(b)). This loading configuration was chosen because it Figure 6 shows the load versus maximum deflection
enabled two tests to be performed on each specimen. Specif- curves for RC beams with transverse steel reinforcement.
ically, while one end zone was being tested, the other end It reveals that each of the specimens in Series HR and MR
zone was overhung and unstressed. The load was applied at exhibited an overall stiffness relatively similar to that of the
a distance a = 3d from the nearest support, a configuration other beams. The maximum load at failure and the maximum
which was representative of a slender beam. deflection attained at the loading point for each specimen
are provided in Table 1. Specimen HR-ST-HF reached the
Instrumentation highest maximum load at failure. Meanwhile, the HR-ST-LF
The measuring equipment used in this research study was and HR-SH specimens exhibited a slightly higher deflec-
carefully designed to meet the objectives of this study. The tion at the loading point than the other strengthened and
vertical displacement was measured at the position under the unstrengthened specimens (Table 1). It should be mentioned
applied load and at midspan, using a linear variable differen- that in Table 1, the shear contributions of concrete and steel
tial transformer (LVDT) with a 150 mm (5-7/8 in.) stroke. were calculated based on the measured experimental results
Different types of strain gauges were installed on the longi- for the control beams.
tudinal reinforcement, on the steel stirrups, and embedded in
the concrete to measure the strains experienced by the various Strain analysis
materials as the loading increased and to monitor thereby This part of the study investigated the behavior of CFRP
the yielding of the steel. The strain gauges on the stirrups and transverse steel during loading of the specimens. As
were installed along the anticipated plane of the major shear mentioned previously, extensive instrumentation for strain
crack. Displacement sensors, also known as crack gauges, monitoring was carefully planned and implemented to
were used to measure the strains experienced by the CFRP provide the data needed to gain a better understanding of
strips and sheets (Fig. 2 and 3). These gauges were fixed the effect of transverse steel on the contribution of FRP to
vertically onto the lateral faces of the specimens at the same

ACI Structural Journal/March-April 2014 355


Fig. 3—Location of strain gauges on transverse steel reinforcement and FRP sheets/strips. Positions of strain gauges on FRP
sheets/strips are identical in specimens strengthened with similar FRP strengthening configurations.
the shear resistance of RC beams retrofitted in shear with maximum strain values for Specimens NR-SH and MR-SH
EB FRP. were 17 and 27%, respectively, of the ultimate strain.
CFRP strain—The distribution of the maximum strains 4. For specimens strengthened with FRP strips in Series
attained in the CFRP is shown in Fig. 7 for all strengthened NR (no steel stirrups), the maximum measured FRP strain
test specimens. It should be noted that these strain values are values were approximately equal. This is also true for the
the maximum measured values, but not necessarily the abso- specimens strengthened with FRP strips in the HR series.
lute maximum values, experienced by the CFRP U-jackets. 5. For all specimens strengthened with FRP strips in both
The two values may differ in cases where the strain gauges Series NR and HR, the maximum FRP strain was greater than
did not intercept the main cracks. From Fig. 7, the following 5000 με. It may be of interest to note that ACI 440.2R-08
observations can be made: (ACI Committee 440 2008) limits the maximum design FRP
1. All curves presented in these figures show that the strain value to 4000 με. On the basis of the results achieved
CFRP did not contribute to load-carrying capacity in the in this study, this limit appears to be conservative for RC
initial stage of loading. beams strengthened with FRP strips.
2. For specimens of Series NR, the measured strains were Transverse steel reinforcement strain—Figure 8 shows the
greater for specimens strengthened with FRP strips than for measured strain in the transverse steel reinforcement for the
similar beams strengthened with FRP sheets. test specimens with internal steel reinforcement. The vertical
3. For the beams strengthened with a layer of FRP sheet line identifies the strains corresponding to the yielding of the
(NR-SH, MR-SH, and HR-SH), the maximum strain in the transverse steel, as obtained by tests (εy = 3250 με). From
CFRP increased as the amount of transverse reinforcement Fig. 8, the following observations can be made:
was increased. In fact, for Specimen HR-SH, the maximum 1. Like the CFRP, the steel stirrups did not contribute to
strain attained by the FRP sheet was approximately 48% load-carrying capacity in the initial stage of loading. The
of the ultimate strain value, whereas the corresponding

356 ACI Structural Journal/March-April 2014


Fig. 4—Failure mode and multiline cracking pattern of strengthened specimens: (a) NR-ST-HF; (b) NR-SH; (c) HR-ST-LF;
and (d) HR-ST-HF after failure.

Fig. 5—Load versus deflection at load point: Series NR. Fig. 6—Load versus deflection at load point: Series HR and
(Note: 1 mm = 0.0394 in.; 1 kN = 0.225 kip.) MR. (Note: 1 mm = 0.0394 in.; 1 kN = 0.225 kip.)
transverse steel contribution to shear resistance started after verse steel yields at ultimate strain for RC beams strength-
the formation of diagonal cracking initiated. ened in shear with EB FRP.
2. All monitored stirrups were significantly strained. This Figure 8 shows that addition of EB FRP did not result in a
is also reflected by the cracking pattern observed in the decrease of transverse steel strain. For all the specimens with
beams with transverse steel (Fig. 4(a) through (d)). transverse steel, the steel yielded well before the RC beam
3. Yielding of transverse steel was observed in all cases. reached ultimate failure. Therefore, it can be concluded that
This observation is in good agreement with existing code at the ultimate state the contribution of internal steel stir-
specifications and guidelines, which assume that the trans- rups to shear resistance was not affected by the addition of

ACI Structural Journal/March-April 2014 357


Fig. 7—Load versus strain in FRP for all strengthened spec- Fig. 8—Load versus strain in transverse steel: Series HR
imens. (Note: 1 kN = 0.225 kip.) and MR. (Note: 1 kN = 0.225 kip.)
externally bonded FRP. It follows that the shear contribu- The experimental contribution of transverse steel Vs is
tion of internal steel reinforcement Vs should be calculated calculated as the sum of the contributions corresponding
using the same formula for both FRP-strengthened and to the stirrups crossing the plane of rupture using the
unstrengthened RC beams, which confirms the assumptions following equation
of the design guidelines (ACI 440.2R-08; CSA S806-02;
Oehlers et al. 2008). Vs = As Es ∑ e s ,i (2)
These results are not in agreement with those based on
finite-element simulations reported by Chen et al. (2010, where As is the section area of one stirrup; Es is the elastic
2012); these researchers found that for RC beams strength- modulus of the transverse steel; and εs,I (≤εy) is the measured
ened in shear with FRP, the internal steel stirrups did not strain in stirrup i in the failure zone, where εy is the yield
reach the yield point. Based on their finite-element model, strain of the stirrups.
they concluded that the yield strength of the internal steel The experimental contribution of FRP Vf can be calculated
stirrups in such strengthened RC beams cannot be fully used. as follows
The models by Chen et al. (2010, 2012) were originally
generated based on a single crack failure pattern assump- V f = 2 E f t f ∑ (wi e f ,i ) (3)
tion. Single crack failure pattern was adopted by most design
models to simplify the calculation and design of strength- where Ef is the elastic modulus of the CFRP; tf is the thick-
ening FRP sheets and strips. Experimental observations, ness of the CFRP; εf,i is the measured strain in the CFRP
however, clearly show that for RC beams strengthened with corresponding to instrumented section i in the failure zone;
EB FRP, the cracking pattern on the FRP-concrete interface and wi is the tributary width of the strengthening FRP strips
is rather distributed (Mofidi and Chaallal 2011a). Eventually, intercepted by the major shear crack, where the CFRP strain
the distributed cracks at the concrete cover merge together at εf,i is assumed constant. The CFRP strengthening width
the concrete core to form one major shear crack at ultimate. represents the portion of the CFRP that effectively contrib-
Therefore, it is believed that assuming a single crack pattern utes to shear resistance.
is overly simplistic when considering such a precise finite Figure 9 shows the experimental evolution under
element modeling tool. Considering the fact that the crack increasing load of the contributions to the shear resistance
width plays a governing role in the Chen et al. (2010, 2012) of the two components (Vs and Vf) for Specimens HR-NF,
models, the discrepancies between the results produced by HR-ST-LF, HR-ST-HF, and HR-SH. The specimens shown
Chen et al. (2010, 2012) models and the experimental results in this figure had the same degree of transverse steel rein-
are to be expected. forcement (highly reinforced), but were strengthened using
different amounts of externally bonded FRP strips and sheet.
Shear resistance under increasing load The behavior of the transverse steel under increasing load
In accordance with most codes and standard guidelines, exhibited a similar pattern for both the unstrengthened beam
the nominal shear resistance Vn can be expressed as follows (HR-NF) and the beams retrofitted with different amounts
of EB FRP strips and sheet (HR-ST-LF, HR-ST-HF, and
Vn = Vc + Vs + Vf (1) HR-SH). This result indicates that strengthening of RC
beams with EB FRP does not alter the behavior of internal

358 ACI Structural Journal/March-April 2014


Table 2—Coefficient of determination (R2) between
values of Vf as calculated by each of the models
and experimental values of Vf
Vf cal
Vf cal by Mofidi
by ACI Vf cal and
440.2R by HB 305 Chaallal
Specimen Vfexp (2008) (2008) (2011a)
NR-ST-LF 53.3 20.5 32.5 43.7
NR-ST-HF 69.3 28.6 39.8 37.9
NR-SH 38.7 40.9 45.9 35.5
HR-ST-LF 14.5 20.5 32.5 7.3
HR-ST-HF 21.7 28.6 39.8 8.3
HR-SH 18.4 40.9 45.9 10.2
MR-SH 27.3 44.1 50.7 14.4
2
R 0.04 0.03 0.81
Fig. 9—Transverse steel and FRP contributions under
increasing load for Specimens HR-NF, HR-ST-LF, Notes: Vfexp is experimental shear resistance due to FRP; Vfcal is calculated shear
HR-ST-HF, and HR-SH. (Note: 1 mm = 0.0394 in.; 1 kN = resistance due to FRP (not factored); 1 kN = 0.225 kip.

0.225 kip.)
RC beams. This result confirms that increasing the amount
of transverse steel leads to a reduction in the contribution of
the FRP during loading and at the ultimate state.

COMPARISON OF TEST RESULTS WITH


SHEAR DESIGN EQUATIONS
The shear resistance due to CFRP as obtained by tests
Vfexp is compared in Table 2 to the nominal shear resistance
Vfcal predicted by ACI 440.2R (2008), HB 305 (Oehlers et al.
2008), and Mofidi and Chaallal (2011a).
Mofidi and Chaallal (2011a) proposed a model for calcu-
lating the contribution of FRP to shear resistance, taking into
consideration the attenuating effect of transverse steel as
well as of the cracking pattern on the EB FRP contribution in
shear. Based on their study, it was determined that the pres-
ence of transverse steel favors the formation of a multi-line
shear-cracking pattern in the RC beam, which decreases the
anchorage length of the FRP fibers and hence the available
effective width of FRP wfe and bonding area between the
Fig. 10—Transverse steel and FRP contributions under
FRP and the concrete. In the calculation of wfe, it is assumed
increasing load for Specimens HR-NF, MR-SH, and HR-SH.
that the cracking pattern of the RC beam becomes more
(Note: 1 kN = 0.225 kip.)
propagated with the increase in the amount of internal steel
transverse steel. It also reveals that addition of EB FRP and external FRP shear reinforcement as measured by their
does not attenuate the shear contribution of the transverse respective rigidities. On the other hand, the cracking pattern
steel reinforcement. influences the anchorage length of the fibers. As the cracking
Figure 10 shows the experimental progression of the pattern becomes more propagated, fewer fibers will provide
shear contributions of the two components (Vs and Vf) under the minimum effective anchorage length. Therefore, the
increasing load for Specimens HR-NF, MR-SH, and HR-SH. effective width, that is the width of the fibers long enough
The strengthened specimens illustrated in this figure were to attain the effective anchorage length, decreases. Using a
both retrofitted with one layer of CFRP sheet, but reinforced computational analysis based on the available test results
with different amounts of transverse steel reinforcement. in the literature (Mofidi and Chaallal 2011a), the effective
The behavior of the FRP under increasing load followed width is defined as a function of the sum of the rigidities
different patterns for the strengthened beams depending of transverse steel reinforcement and that of transverse FRP
on the amount of internal transverse steel reinforcement sheets (Eq. (9) and (10)).
(MR-SH and HR-SH). This clearly shows that the behavior
of EB FRP depends on the amount of transverse steel used in 0.6
w fe = × df for U-jacket (4)
ρ f ⋅ E f + ρs ⋅ E s

ACI Structural Journal/March-April 2014 359


0.43
w fe = × d f for side bonded (5)
ρ f ⋅ E f + ρs ⋅ E s

With wfe defined, the cracking modification factor can then


be calculated as kc = wfe/df—that is

w fe 0.6
kc = = for U-jackets (6)
df ρ f ⋅ E f + ρs ⋅ E s

w fe 0.43
kc = = for side bonded (7)
df ρ f ⋅ E f + ρs ⋅ E s

The effect of kL for beams with an anchorage length less


than the effective length and that of kf counting for the wf /sf
ratio of the FRP strips are considered in the equation for
effective strain, as follows
Fig. 11—FRP contribution: values as calculated by ACI
k ⋅k ⋅k ⋅τ ⋅L fc′ 440.2R-08, HB 305-08, and Mofidi and Chaallal (2011a)
e fe = c L f eff e = 0.31kc ⋅ kL ⋅ k f ≤ e fu (8)
tf ⋅ Ef tf Ef models versus experimental values. (Note: 1 kN = 0.225 kip.)

The Mofidi and Chaallal (2011a) model justifies prema-
It should be noted that kL and kf can be calculated using ture FRP debonding in RC beams with internal transverse
Eq. (9) and (11), as follows steel reinforcement compared with beams with no transverse
steel and explains the superior gain achieved due to FRP in
 1 if l ≥ 1  Lmax beams with few or no steel stirrups compared with beams
kL =   l= (9)
( 2 − l ) ⋅ l if l < 1 Le with moderate to high amounts of internal transverse steel.

To test the correlation between the experimental results
and the predicted results by the models, the best fit of the
where
nominal predicted results versus the experimental results
was considered. The following assumptions were made
 df  when calculating the experimental results: 1) the shear resis-
 for U-jackets 
 sin b  tance due to concrete was assumed constant for beams with
Lmax =  (10) or without transverse steel reinforcement; 2) the shear resis-
 df for side plates 
tance due to concrete was assumed constant for both retro-
 2 sin b 
fitted and unstrengthened specimens; and 3) the contribution
of the transverse steel was assumed constant for both retro-
fitted and unstrengthened specimens.
 wf 1
2
For specimens with no transverse steel strengthened using
1+  −  FRP strips (NR-ST-LF and NR-ST-HF), all three models
 sf 2 underestimated the shear resistance due to FRP. This effect
kf = 2 (11)
 wf 1 was more significant when using ACI 440.2R-08 (ACI
1−  −  Committee 440 2008), where, for example, for NR-ST-HF,
 sf 2 the shear resistance predicted was 28 kN (6.3 kip), compared

with 69.3 kN (15.6 kip) obtained by test. On one hand, for
In addition, Vf can be calculated as a function of εfe using the specimen with no transverse steel strengthened using an
the following equation that accounts for the effect of the FRP sheet (NR-SH), the ACI 440.2R-08 (ACI Committee
cracking angle θ (the cracking angle can be taken equal to 440 2008) and HB 305-08 (Oehlers et al. 2008) models
45 degrees for simplicity) slightly overestimated the shear resistance due to FRP.
On the other hand, the Mofidi and Chaallal (2011a) model
2t f ⋅ w f ⋅ e fe ⋅ E f ⋅ (cot θ + cot a ) ⋅ sin a ⋅ d f slightly underestimated the FRP contribution to shear resis-
Vf = tance (Table 2). For all strengthened specimens with trans-
sf
(12) verse steel reinforcement, the ACI 440.2R-08 and HB 305-08
models provided unconservative predictions and therefore
= ρ f ⋅ E f ⋅ e fe ⋅ b ⋅ d f ⋅ (cot θ + cot a ) ⋅ sin a overestimated results (Table 2). In contrast, results produced

by Mofidi and Chaallal (2011a) for specimens strengthened
with transverse steel reinforcement correlate fairly well with

360 ACI Structural Journal/March-April 2014


the test results. Figure 11 shows that the Mofidi and Chaallal strain (that is, 4000 με) seems conservative for RC beams
(2011a) model predicted the experimental shear contribution strengthened in shear with FRP strips.
of FRP (R2 = 0.81) with a high level of accuracy. The ACI
440.2R-08 and HB 305-08 models produced low coefficients AUTHOR BIOS
of determination (0.04 and 0.03, respectively). In general, Amir Mofidi is a Postdoctoral Fellow in the Department of Civil Engi-
neering and Applied Mechanics of McGill University, Montreal, QC,
current design guidelines models (including ACI 440.2R-08 Canada. He received his PhD from University of Quebec, École de Tech-
and HB 305-08) fail to consider the effect of the transverse nologie Supérieure, Montreal, QC, Canada. His research interests include
steel (Mofidi and Chaallal 2011a,b). Therefore, they may the use of fiber-reinforced polymer composites for strengthening and retro-
fitting of concrete structures.
predict conservative results for beams without transverse
steel reinforcement. In contrast, current design guidelines Omar Chaallal, FACI, is a Professor of construction engineering,
models may overestimate the shear contribution of FRP University of Quebec, École de Technologie Supérieure. He is a member
of ACI Committee 440, Fiber-Reinforced Polymer Reinforcement. His
for the specimens with transverse steel reinforcement and research interests include experimental and analytical research on the use
hence, the shear resistance. Such unconservative results are of fiber-reinforced polymer composites for reinforcement and repair of
exemplified in the results predicted by ACI 440.2R-08 and concrete structures.
HB 305-08 for Specimens HR-ST-LF, HR-ST-HF, HR-SH,
ACKNOWLEDGMENTS
MR-SH in the current study. The presence of the transverse The financial support of the National Science and Engineering Research
steel has a significant effect on the shear resistance of RC Council of Canada (NSERC), the Fonds québécois de la recherche sur la
beams strengthened with FRP, and therefore, should ulti- nature et les technologies (FQRNT), and the Ministère des Transports du
Québec (MTQ) is gratefully acknowledged. The efficient collaboration of
mately be considered in design models. J. Lescelleur (Senior Technician) and J. M. Rios (Technician) at ÉTS in
conducting the tests is acknowledged.
CONCLUSIONS
This paper presents the results of an experimental inves- REFERENCES
tigation involving 10 tests on RC T-beams strengthened ACI Committee 440, 2008, “Guide for the Design and Construction of
Externally Bonded FRP Systems for Strengthening Concrete Structures
in shear with EB FRP strips and sheets. The effects of (440.2R-08),” American Concrete Institute, Farmington Hills, MI, 76 pp.
the following parameters were studied: 1) the CFRP ratio Bousselham, A., and Chaallal, O., 2004, “Shear-Strengthening Rein-
(that is, the spacing of the CFRP strips); 2) the presence or forced Concrete Beams with Fiber-Reinforced Polymer: Assessment of
Influencing Parameters and Required Research,” ACI Structural Journal,
absence of transverse steel; 3) the transverse steel ratio (that V. 101, No. 2, Mar.-Apr., pp. 219-227.
is, the spacing between the stirrups); and 4) the use of CFRP Bousselham, A., and Chaallal, O., 2008, “Mechanisms of Shear Resis-
strips versus CFRP sheets. The following conclusions can tance of Concrete Beams Strengthened in Shear with Externally Bonded
FRP,” Journal of Composites for Construction, ASCE, V. 12, No. 5,
be drawn: pp. 499-512.
1. The addition of internal transverse steel reinforcement Chaallal, O.; Mofidi, A.; Benmokrane, B.; and Neale, K., 2011,
resulted in a significant decrease in the gain due to FRP for “Embedded Through-Section FRP Rod Method for Shear Strengthening
of RC Beams: Performance and Comparison with Existing Techniques,”
all the strengthened specimens. Journal of Composites for Construction, ASCE, V. 15, No. 3, pp. 374-383.
2. For all test specimens with transverse steel reinforce- Chaallal, O.; Nollet, M. J.; and Perraton, D., 1998, “Strengthening
ment, the steel yielded before the specimen failed. The of Reinforced Concrete Beams with Externally Bonded Fiber-Rein-
forced-Plastic Plates: Design Guidelines for Shear and Flexure,” Canadian
presence of externally bonded FRP for shear retrofit did Journal of Civil Engineering, V. 25, No. 4, pp. 692-704.
not cause a significant decrease in transverse steel strain. Chen, G. M.; Teng, J. G.; and Chen, J. F., 2012, “Shear Strength Model
Overall, the contribution of steel stirrups to shear resistance for FRP-Strengthened RC Beams with Adverse FRP-Steel Interaction,”
Journal of Composites for Construction, ASCE, DOI: 10.1061/(ASCE)
was not adversely affected by the addition of FRP. CC.1943-5614.0000313.
3. Comparison of the resistance predicted by the Chen, G. M.; Teng, J. G.; Chen, J. F.; and Rosenboom, O. A., 2010,
ACI 440.2R-08 (ACI Committee 440 2008), HB 305-08 “Interaction between Steel Stirrups and Shear-Strengthening FRP Strips in
RC Beams,” Journal of Composites for Construction, ASCE, V. 14, No. 5,
(Oehlers et al. 2008), and Mofidi and Chaallal (2011a) pp. 498-509.
models with test results showed that the guidelines failed to Khalifa, A.; Gold, W. J.; Nanni, A.; and Aziz, A., 1998, “Contribution
capture the influence of transverse steel on the shear contri- of Externally Bonded FRP to Shear Capacity of RC Flexural Members,”
Journal of Composites for Construction, V. 2, No. 4, pp. 195-203.
bution of FRP. The model proposed by Mofidi and Chaallal Mofidi, A., and Chaallal, O., 2011a, “Shear Strengthening of RC Beams
(2011a) showed a better correlation with experimental with Epoxy-Bonded FRP—Influencing Factors and Conceptual Debonding
results than the guidelines mentioned. Model,” Journal of Composites for Construction, ASCE, V. 15, No. 1,
pp. 62-74.
4. The maximum measured strain values in CFRP strips, Mofidi, A., and Chaallal, O., 2011b, “Shear Strengthening of RC Beams
and hence the gain in shear strength due to CFRP strips, with Externally Bonded FRP Composites: Effect of Strip-Width to Strip-
were significantly greater than for CFRP continuous sheets. Spacing Ratio,” Journal of Composites for Construction, ASCE, V. 15,
No. 5, pp. 732-742.
In addition, the maximum deflection was slightly greater for Oehlers, D. J.; Seracino, R.; and Smith, S. T., 2008, Design Guideline
beams retrofitted with CFRP strips than for beams strength- for RC Structures Retrofitted with FRP and Metal Plates: Beams and Slabs,
ened with continuous CFRP sheets. HB 305-2008, Standards Australia, Sydney, Australia, 73 pp.
Triantafillou, T. C., 1998, “Shear Strengthening of Reinforced Concrete
5. In all the specimens strengthened with FRP strips, the Beams Using Epoxy-Bonded FRP Composites,” ACI Structural Journal,
maximum attained FRP strain was greater than 5000 με. It V. 95, No. 2, Mar.-Apr. pp. 107-115.
follows that the ACI 440.2R-08 limit for maximum FRP Uji, K., 1992, “Improving Shear Capacity of Existing Reinforced
Concrete Members by Applying Carbon Fiber Sheets,” Transactions of the
Japan Concrete Institute, V. 14, pp. 253-266.

ACI Structural Journal/March-April 2014 361


NOTES:

362 ACI Structural Journal/March-April 2014


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 111-S32

Punching of Reinforced Concrete Flat Slabs with Double-


Headed Shear Reinforcement
by Maurício P. Ferreira, Guilherme S. Melo, Paul E. Regan, and Robert L. Vollum
Twelve slabs, 11 of which contained double-headed studs as
shear reinforcement, were tested supported by central column
and loaded concentrically. Their behavior is described in terms of
deflections, rotations, strains of the concrete close to the column,
strains of the flexural reinforcement across the slab width, and
strains of the studs. All failures were by punching, in most cases
within the shear reinforced region. The treatments of punching
resistance in ACI 318, Eurocode 2 (EC2), and the critical shear
crack theory (CSCT) are described, and their predictions are
compared with the results of the present tests and 39 others from
the literature. The accuracy of predictions improves from ACI 318
to EC2 to CSCT—that is, with increasing complexity. However, the
CSCT assumptions about behavior are not well supported by the
experimental observations.

Keywords: codes; flat slabs; punching; shear studs.

INTRODUCTION
There is no generally accepted theoretical treatment of
punching, and design is based on empirical methods given in
codes of practice. While there is similarity between them in
terms of general approach, there are considerable differences
in their assumptions and the resulting equations, which leads
to uncertainties about their reliability.
A further cause of uncertainty is the wide variety of types
of shear reinforcement, such as stirrups of various forms,
bent-up bars, welded fabric, and stud systems. Comparisons
of design equations with the results of tests using different
types of shear reinforcement can result in a wide scatter,
while comparisons of slabs with only one type are often
limited by the restricted data available.
This paper presents the results of tests1 of slabs with Fig. 1—Test arrangements. (Note: Dimensions in mm;
double-headed studs as shear reinforcement, followed by 1 mm = 0.0394 in.)
a short review of the design methods of ACI 318,2 Euro-
code 2 (EC2),3 and the critical shear crack theory (CSCT) of experimental and calculated strengths for the present tests
Muttoni et al.,4,5 which is the basis of the punching clauses and others are presented to evaluate the accuracies of
of the fib Model Code 2010 draft.6 The results of the present the methods.
tests and of others on slabs with double-headed shear rein-
forcement are then compared with the three design methods. EXPERIMENTAL PROGRAM
Twelve tests were made at the University of Brasilia. The
RESEARCH SIGNIFICANCE specimens were square slabs 2.5 x 2.5 m (8.2 x 8.2 ft) on plan
There are considerable differences between the design and 180 mm (7.1 in.) thick supported centrally by circular or
methods for punching in ACI 318, EC2, and the CSCT. square columns (Type C and S slabs, respectively). Equal
The primary objective of the experimental study described downward loads were applied at eight points close to the
in this paper was to assess the realism of the assumptions slab edges, as shown in Fig.1.
underlying these design methods. The principal variables in
the test series were the sizes and spacings of the studs, and ACI Structural Journal, V. 111, No. 2, March-April 2014.
MS No. S-2012-119, doi:10.14359.51686535, was received April 4, 2012, and
the size and shape of the columns. Extensive measurements reviewed under Institute publication policies. Copyright © 2014, American Concrete
were made of slab rotations and strains in the concrete, and Institute. All rights reserved, including the making of copies unless permission is
obtained from the copyright proprietors. Pertinent discussion including author’s
flexural and shear reinforcement. Comparisons between closure, if any, will be published ten months from this journal’s date if the discussion
is received within four months of the paper’s print publication.

ACI Structural Journal/March-April 2014 363


The main variables were the shape and size of the column,
the amount and distribution of the shear reinforcement, and
some details of the main reinforcement.
The concrete was made with ordinary portland cement,
natural sand, and crushed limestone aggregate with a
maximum size of 9.5 mm (3/8 in.). The concrete strength
was determined from 100 x 200 mm (4 x 8 in.) control cylin-
ders that were tested at the same time as the slabs.
The arrangement of flexural reinforcement was basically
the same in all but two of the specimens (Slabs C5 and C6).
The general arrangement of the upper tension reinforcement
was 16 mm (No. 5) bars with fy = 540 MPa (78 ksi) and
Es = 213 GPa (30,893 ksi) at spacings of 100 mm (4 in.)
in the outer layer and 90 mm (3.54 in.) in the inner layer,
providing almost equal flexural resistances in two directions.
The bottom reinforcement was 8 mm (0.315 in.) bars posi-
tioned directly below alternate top bars. At the edges, each
top bar was lapped with a 12.5 mm (No. 4) hair-pin shaped
bar with 500 mm (20 in.) horizontal legs. Only minor adjust-
ments to this arrangement were needed to avoid clashes with
shear reinforcement.
In Slab C5, the tension reinforcement in the central parts Fig. 2—Flexural reinforcement of Slab C5. (Note: Dimensions
of the widths was increased to 20 mm (No. 6) bars with fy = in mm; 1 mm = 0.0394 in.)
544 MPa (79 ksi) and Es = 208 GPa (30,168 ksi) and that in
the outer parts was decreased, to obtain a higher reinforce- floor. An example of the deflected profiles is shown in
ment ratio near the column without significantly altering the Fig. 4, where it can be seen that segments of the slab rotated
flexural capacity. The details of this slab are shown in Fig.2. about axes at or very close to the column face, and the top
As the failures of some of slabs appeared to be influenced surfaces remained more or less straight on radial lines. The
by crushing of the soffit near the column, Slab C6 was displacements of the slab very close to the column, visible in
provided with compression reinforcement comprised of four Fig. 4, were likely due mostly to movements in the support
16.0 mm (No. 5) bars through the column in each direction, of the column, and were not deflections of the slab relative
and 12.5 mm (No. 4) bars below all the top bars in the rest to the column.
of the width. Figure 5 shows envelopes of the experimental load-
The shear reinforcement was double-headed studs made rotation relationships, and the theoretical ones according
of deformed 10 mm (No. 3) bars with fyw = 535 MPa (78 ksi) to CSCT. The experimental rotations plotted are the aver-
and Es = 211 GPa (30,603 ksi), or 12.5 mm (No. 4) bars with ages of values determined from deflections, measured on
fyw = 518 MPa (75 ksi) and Es = 204 GPa (29,588 ksi). The the centerline of the slabs at distances 274 and 1049 mm
heads, with diameters three times the bar size, were welded (10.8 and 41.3 in.) from the slab center in the North, South,
to the shanks, and the completed studs were spot-welded to East, and West directions. The CSCT values have been
nonstructural carrier rails, which were 10 mm (3/8 in.) wide calculated from Eq. (13) and (14). The correlations between
and 3.2 mm (1/8 in.) thick. The shear reinforcement was experimental and calculated results are good.
positioned from above, with the carrier rails sitting on the
upper tension bars either directly or via cross rails. Strains of concrete
Tests of studs, in which the loading was applied via the Strain gauges were used to measure the strains of the
heads, showed that the welds between the heads and shanks bottom surfaces of the concrete close to the columns. The
were able to develop the full strengths of the bars with general responses were similar to those observed by others.
ductile failures away from the welds. At low loads, the compression strains in both directions
In all but one of the slabs, the lines of studs ran outward were similar, and increased with increasing load. As loading
from the columns along equally spaced radial lines (radial continued, the radial strains stabilized and then decreased,
arrangement). The exception was Slab C4, where a cruci- sometimes becoming tensile before failure. The tangential
form arrangement was used. Typical details are shown in compressions increased at progressively higher rates.
Fig. 3. Table 1 summarizes the characteristics of all slabs. At circular columns, the maximum compression strains
were from 2.5 to 2.8‰, where the gauges were 20 mm
TEST RESULTS (0.8 in.) from the column faces, and 3.1 to 3.2‰, where the
gauges were 40 mm (1.6 in.) from the faces. In the slabs
Deflections and rotations with square columns, the maximum compressions recorded
Deflections of the top surfaces of the slabs were measured were lower, but this was probably due to their being on the
along their centerlines by dial gauges mounted from frames slab centerlines, while the greatest strains were likely at the
spanning over the slabs and supported on the laboratory corners of the columns.

364 ACI Structural Journal/March-April 2014


Table 1—Characteristics of test slabs
Shear reinforcement
Slab No. *
Column size , mm (in.) d, mm (in.) ρ,%

fc, MPa (ksi) Studs ‡
so, mm (in.) sr, mm (in.)
C1 270 (10.6) 143 (5.6) 1.48 47.8 (6.9) 10 φ10.0 x 6 70 (2.8) 100 (3.9)
C2 360 (14.2) 140 (5.5) 1.52 46.9 (6.8) 10 φ10.0 x 6 70 (2.8) 100 (3.9)
C3 450 (17.7) 142 (5.6) 1.49 48.9 (7.1) 10 φ10.0 x 6 70 (2.8) 100 (3.9)
C4 360 (14.2) 140 (5.5) 1.52 47.9 (6.9) 12 φ10.0 x 6 70 (2.8) 100 (3.9)
C5 360 (14.2) 140 (5.5) 2.00 49.7 (7.2) 10 φ10.0 x 6 70 (2.8) 100 (3.9)
C6 360 (14.2) 143 (5.6) 1.48 48.6 (7.0) 10 φ10.0 x 6 70 (2.8) 100 (3.9)
C7 360 (14.2) 144 (5.7) 1.47 49.0 (7.1) 10 φ10.0 x 7 55 (2.2) 80 (3.1)
C8 360 (14.2) 144 (5.7) 1.47 48.1 (7.0) 12 φ10.0 x 6 70 (2.8) 100 (3.9)
S1 300 (11.8) 145 (5.7) 1.46 48.3 (7.0) 12 φ10.0 x 2 70 (2.8) 100 (3.9)
S2 300 (11.8) 143 (5.6) 1.48 49.4 (7.2) 12 φ10.0 x 4 70 (2.8) 100 (3.9)
S5 300 (11.8) 143 (5.6) 1.48 50.5 (7.3) — — —
S7 300 (11.8) 143 (5.6) 1.48 48.9 (7.1) 12 φ12.5 x 4 70 (2.8) 100 (3.9)
*
Diameter in Series C, side length in Series S.

Calculated as ρx ρy . In all slabs except C5, reinforcement distributed uniformly across widths. For C5, 2.00% is ratio in central (c + 6d), and 1.56% is ratio for full width.

Number of studs per perimeter, stud size (mm) by number of perimeters.

Fig. 3—Shear reinforcement. (Note: Dimensions in mm;


1 mm = 0.0394 in.)
The difference in maximum strains at distances of 20 and
40 mm (0.8 and 1.6 in.) points to restraint from the columns,
and the strains 40 mm (1.6 in.) from the columns were
probably high enough to indicate distress of the concrete
due to tangential stresses, except in the one slab without
shear reinforcement. Fig. 4—Load-displacement of Slab C3. (Note: 1 mm =
0.0394 in.)

ACI Structural Journal/March-April 2014 365


Fig. 5—Load-rotation behavior of tested slabs. (Note:
1 kN = 0.2248 kip.) Fig. 6—Strains of concrete at soffits of Slabs C2 and C6.
(Note: 1 kN = 0.2248 kip.)
Figure 6 shows the developments of strains measured in
Slabs C2 and C6, the former having only nominal bottom The strains were measured in the inner three rings of shear
steel, and the latter being the slab with considerable compres- reinforcement in the slabs with circular columns, and at all
sion reinforcement. In C6, the maximum strain of 3.2‰ was perimeters for the slabs with square columns. The average
reached at a load equal to the failure load of C2. There- stresses (strain × Es) in the shear reinforcement are summa-
after, the strain in C6 decreased as the load was increased to rized in Table 2. Typical profiles of stud stresses along radial
failure. This suggests that the 12% higher ultimate strength lines are shown in Fig. 8. The stud stresses summarized in
was achieved with the compression reinforcement locally Table 2 are thought to be reasonably close to the maximum
taking over the function of the failing concrete. stresses in the studs because the studs were short. The differ-
ence between the measured and maximum stress depends on
Strains of flexural reinforcement the product of the bond stress and the distance from the shear
Strains of the flexural tension reinforcement were crack to the midheight of the studs. Allowing for bond along
measured by pairs of strain gauges at opposite ends of diam- lengths between cracks and strain gauges, it appears that
eters of the upper bars, at a section just outside the column. the first perimeter of shear reinforcement is likely to have
The resulting profiles of tangential strains for Slabs C1 yielded in all the tests except C6, C7, and S7.
through 4 and C8 are shown in Fig. 7. Strains beyond yield
were recorded in considerable parts of the slab widths, but Ultimate loads and modes of failure
the yielding never reached the slab edge, that is, a yield line All of the slabs failed by punching, and Table 2 gives the
was never developed. ultimate loads and summarizes data from relevant strain
measurements at or close to failure. The slabs with shear
Strains of shear reinforcement reinforcement failed inside the shear reinforced areas in all
Strains were measured at the midheight of the shear cases but S1, where there were only two perimeters of studs,
studs in four lines of the shear reinforcement in all slabs. and S7, where the diameter of the studs was 12.5 mm (No. 4).

366 ACI Structural Journal/March-April 2014


Fig. 7—Strains of flexural reinforcement. (Note: 1 mm =
0.0394 in.; 1 kN = 0.2248 kip.) Fig. 8—Average stud stresses at Perimeters 1 to 3. (Note:
1 kN = 0.2248 kip; 1 MPa = 0.1450 kip.)

ACI Structural Journal/March-April 2014 367


Specimens C2, C5, and C6 (Table 1) were similar apart The specimens were saw-cut half width in two orthogonal
from the detailing of the flexural reinforcement. Comparison directions to reveal the failure surfaces. For the inside fail-
of the shear strengths of C2 and C5 shows that the punching ures, most of the surfaces ran from the soffit at the column
resistance was increased by approximately 15% by concen- face to reach the level of the top steel at the second perim-
trating 60% of the flexural reinforcement into a 1 m (39 in.) eter of studs, but there were exceptions. In both sections of
wide band centered on the column whilst maintaining the Slab C1 and one of Slab C2, the failure surfaces crossed the
same flexural capacity across the slab width as in the other inner studs very close to their upper heads. In Slabs C5, C6,
tests. The shear strength of C6 was increased by approxi- and S2, they reached to the level of the top bars at the third or
mately 12% relative to C2 through the provision of addi- fourth perimeter of studs. In C7, there were multiple cracks
tional compression steel. reaching the main steel from the third to the fifth perimeter
in one section, while in the perpendicular direction, the
surface ran just above the lower heads of the studs out to the
fourth layer and reached the top steel at the sixth layer. In the
outside failures, the surface was entirely outside the studs in
S1, but did cross them just above their lower heads in S7. In
some slabs, most notably C6, there was spalling of the slab
around the column that commenced before failure.

METHODS OF CALCULATION
All three approaches considered herein take the punching
strength of a slab with shear reinforcement as the least of
VR,cs, VR,out, and VR,max, but not less that VR,c, where VR,c is the
resistance of an otherwise similar slab without shear rein-
forcement; VR,cs is the combined resistance of the concrete
and shear reinforcement; VR,out is the resistance from the
concrete alone just outside the shear reinforcement; and
VR,max is the maximum resistance possible for a given column
size, slab effective depth and concrete strength.
These resistances correspond to failures of the types
shown in Fig. 9. The calculations are made for perimeters at
specified distances from supports: uo is the perimeter at the
outline of the support; u1 is the perimeter used in the calcu-
lation of VR,c and VR,cs; and uout is the perimeter used in the
Fig. 9—Types of punching failure. calculation of VR,out.

Table 2—Summary of test results


Average stud stresses‡, MPa (ksi)
Slab No. εc,max , ‰
* †
ry , mm (in.) 1 2 3 Vu§, kN (kip) Failure mode
C1 2.66 450 (17.7) 535 (77.6) 317 (46.0) 137 (19.9) 858 (192.9) In
C2 2.81 550 (21.7) 530 (76.9) 235 (34.1) 121 (17.5) 956 (214.9) In
C3 2.54 625 (24.6) 511 (74.1) 362 (52.5) 189 (27.4) 1077 (242.1) In
||
C4 2.28 770 (30.3) 535 (77.6) 461 (66.8) 297 (43.1) 1122 (252.2) In
C5 3.24 490 (19.3) 504 (73.1) 264 (38.3) 160 (23.2) 1117 (251.1) In
C6 3.20 750 (29.5) 479 (69.5) 421 (61.0) 474 (68.7) 1078 (242.3) In
C7 3.14 540 (21.3) 386 (56.0) 419 (60.8) 167 (24.2) 1110 (249.5) In
C8 3.14 660 (26.0) 535 (77.6) 436 (63.2) 179 (26.0) 1059 (238.1) In
S1 2.37 560 (22.0) 535 (77.6) 473 (68.6) — 1021 (229.5) Out
S2 2.15 570 (22.4) 535 (77.6) 514 (74.5) 216 (31.3) 1127 (253.4) In
S5 1.47 130 (5.1) — — — 779 (175.1) —
S7 2.67 600 (23.6) 238 (34.5) 285 (41.3) 137 (19.9) 1197 (269.1) Out
εc,max is maximum tangential strain of concrete (measured 20 mm from columns in C1 to C4, S1 and S2, and 40 mm from columns in C5 to C8 and S7). For slab Type S, strains
*

measured on centerlines.

ry is radius in which tangential strain > εy.

Averages of Esεs ≤ fyw in Perimeters 1, 2, and 3.
§
Ultimate shear force including self-weights of slabs and loading system.
||
Measured at 0.85Vu.

368 ACI Structural Journal/March-April 2014


Fig. 12—Control perimeters: CSCT.

Fig. 10—Detailing and control perimeters: ACI 318. VR,cs = 0.75VR,c + VR,s (2)

d
VR,s = ⋅ Asw f yw with fyw ≤ 414 MPa (60,000 psi) (3)
sr

VR,out = 1 fc uout d (4)


6

VR,max = 2 fc u1 d if sr ≤ 0.5d (5a)


3

VR,max = 1 fc u1 d if 0.5d ≤ sr ≤ 0.75d (5b)


2

fc is limited to ≤ 69 MPa (10,000 psi) for calculation purposes.


Fig. 11—Detailing and control perimeters: EC2.
EC2-04
The locations and lengths of u1 and uout vary with the
method of calculation. VR,c = 0.18k(100ρfc)1/3u1d (6)
The symbols used for spacings of shear reinforcement are
as follows: so is the distance from column to inner studs; sr k = 1 + 200 / d ≤ 2 (7)
is the radial spacing of studs; and st is the tangential spacing
of studs at a perimeter. The effective depth d is taken as
the average for orthogonal directions, d = (dx + dy)/2. The VR.cs = 0.75VR,c + VR,s (8)
expressions for punching resistances are given below in SI
units (N and mm) without any explicit safety factors. Those d
from ACI 318 are for nominal resistances, and the others VR,s = 1.5 Asw f yw,ef (9)
sr
are for characteristic resistances. The perimeters u1 and uout
and the detailing requirements, in relation to the spacings of
shear reinforcement, are illustrated by Fig. 10, 11, and 12 for fyw,ef = 1.15(250 + 0.25d) ≤ fyw ≤ 600 MPa (87,000 psi) (10)
ACI 318, EC2, and the CSCT, respectively.
VR,out = 0.18k(100ρfc)1/3uout,efd (11)
ACI 318-08
As double-headed studs are not considered explicitly, the  fc 
VR,max = 0.3 fc  1 − u d (12)
equations used herein are those for studs with heads at their  250  o
top ends and bottom anchorages provided by welds to struc-
tural rails
ρ is the ratio of flexural reinforcement calculated as ρx ρy ,
where ρx and ρy are the ratios in orthogonal directions deter-
VR,c = 1 fc u1 d (1)
3 mined for widths equal to those of the column plus 3d to

ACI Structural Journal/March-April 2014 369


VR,s = ΣAswσsi(ψ) ≤ ΣAswfyw (16)

VR,cs = VR,c + VR,s (17)

0.75uout dv fc
VR,out = (18)
1 + 15ψd / (16 + dg )

VR,max = 3VR,c (19)

where σsi is the stress in the i-th perimeter of shear reinforce-


ment which is related to the width of the critical shear crack,
where it crosses the shear reinforcement.5 The summations
ΣAsw and ΣAswσsi(ψ) are for all the shear reinforcement
within a distance d from the column.
Fig. 13—Punching strengths according to CSCT: Slab C1. The CSCT average method is intended to give approxi-
(Note: 1 kN = 0.2248 kip.) mately mean strengths. In it, the stresses in studs at different
either side. ρ ≤ 0.02 for calculation purposes, and the scope distances (≤d) from the column are calculated assuming that
of EC2 is limited to fc ≤ 90 MPa (13,000 psi). the width of the critical shear crack increases linearly, from
zero at the slab soffit to the width corresponding to a slab
Critical shear crack theory rotation ψ and a crack opening angle of 0.5ψ, at the level
In the CSCT, punching resistances are related to the rota- of the tension reinforcement. The stress in a stud is then
tion ψ of the slab, outside a critical crack. Half of this rota- obtained by equating the vertical component of the crack
tion is assumed to occur in the critical shear crack and, as the opening to the elongation of the stud for a given stress at
slab rotates, the concrete component of shear resistance at the crack.
the crack is assumed to decrease, while the component from
the shear reinforcement increases up to yield. The rotation ψ COMPARISONS OF TESTS AND CALCULATIONS
is related to the ratio V/Vflex, where V is the acting shear, and
Vflex is the shear force corresponding to the flexural capacity, General
calculated by yield-line theory. Values of VR,c, VR,cs, VR,out, Experimental strengths from the present tests and from
and VR,max can be determined as shown in Fig. 13, by plotting others reported in the literature have been compared with
the resistances against ψ and finding their intersections with resistances calculated by the three methods described previ-
Eq. (13) ously. The shear reinforcement in the tests by Regan,7 Regan
and Samadian,8 Beutel,9 and Birkle10 was double-headed
3 studs made from either deformed or plain round bars. In the
r fy  V  2 tests by Gomes and Regan,11 it was slices of steel I-beams
ψ = 1.5 s  
d Es  V flex  (13) with the flanges acting as anchorages. The shear reinforce-
ment was positioned radially unless noted as ACI type in
Table A1 in Appendix A.
where rs is the distance from the column center to the line The calculations of punching resistances were made using
of radial contraflexure; and ψ is in radians. For typical the expressions given previously, with their limits generally
punching test specimens, rs is the distance from the column respected. Exceptions to this were as follows.
center to the slab edge. The flexural failure load Vflex is For ACI 318 and EC2, the limits on so/d and sr/d were
approximated by5 given a little tolerance. Values of sr/d up to 0.8 were treated
as acceptable, and for EC2 the lower limit so/d < 0.3 was
 rs  waived with values going down to 0.24. EC2 does not
Vflex = 2πmR  r − r  (14) envisage the use of plain round shear reinforcement, but this
 q c
has been ignored, and lower limits on d for the use of shear
reinforcement were ignored. (The least effective depth in the
where mR is the moment resistance per unit length of yield tests used was 124 mm [4.9 in.] in six slabs by Birkle.10)
line; rq is the radius at which loading is applied; and rc is the The CSCT shear strengths were calculated using slab
radius of the column and for square columns can be taken as rotations calculated with Eq. (13), in which Vflex was calcu-
2c/π, where c is the side length of the column. lated with Eq. (14). The stresses in the shear reinforcement
In the CSCT average method given in Reference 5 were calculated in accordance with the recommendations in
(5). The resulting slab rotations were slightly greater than
0.75u1 d fc the measured slab rotations, as illustrated in Fig. 5. The
(15) predicted shear strengths typically increase by less than 5%
1 + 15ψd / (16 + dg )
VR,c = for the slabs tested in this program if measured rotations are
used instead of calculated rotations. For the CSCT, there

370 ACI Structural Journal/March-April 2014


Table 3—Comparisons with test results for slabs Table 4—Statistics of Vu /Vcalc for slabs with shear
without shear reinforcement reinforcement
Vu/Vcalc No. of ACI 318 EC2 CSCT
Slab No. d, mm (in.) ρ, % ACI 318 EC2 CSCT EC2* tests Mean COV Mean COV Mean COV
Ferreira1 Ferreira1
S5 143 (5.6) 1.48 1.30 1.20 1.24 1.10 11 1.47 0.148 1.20 0.160 1.23 0.108

Gomes and Regan 11 Regan7 and Regan and Samadian8


1 159 (6.3) 1.27 1.16 0.96 1.00 0.89 9 1.56 0.100 1.06 0.087 1.05 0.102
1A 159 (6.3) 1.27 1.20 1.01 1.03 0.92 9
Beutel
Birkle10
6 1.72 0.093 1.34 0.087 1.16 0.101
1 124 (4.9) 1.53 1.30 1.11 1.10 0.96
7 190 (7.5) 1.29 1.12 0.94 1.02 0.93 Gomes and Regan11
10 260 (10.2) 1.10 0.88 0.78 0.86 0.78
10 1.76 0.134 1.28 0.081 1.28 0.083
Mean 1.16 0.99 1.04 0.93 10
Birkle
Coefficient of variation 0.13 0.15 0.12 0.11
9 1.35 0.185 1.09 0.105 1.06 0.050
*
EC2 calculations as for EC2, but with no upper limit on k = 1+√(200/d).
Total
are only a few cases in which there were two perimeters of 45 1.56 0.162 1.19 0.136 1.16 0.121
shear reinforcement within a distance d from the support,
but there are some where a second layer was not much ACI 318 and EC2 are basically empirical, but the CSCT
further out (all of the slabs by Gomes and Regan11 where the claims a rational basis. Unfortunately, its modeling of slab
distances varied from 1.0d to 1.04d and Slabs 2, 3, 9, and 12 deformations is incorrect. The rotation ψ is predicted satis-
by Birkle10 where the distances were from 1.09d to 1.18d). factorily by Eq. (13), but, as can be seen from Fig. 4, it is not
The second perimeter has been included in ΣAsw, where the divided equally into movements at the column face and in a
distance was less than 1.05d. shear crack. In addition, the surfaces at which failure occurs
Details of the individual slabs and the results obtained are not at 45 degrees to the slab plane, but have variable
are given in Appendix A, while Tables 3 and 4 summarize geometries. Refer to the section entitled, “Ultimate loads
the results of the comparisons for slabs without and with and modes of failure.”
shear reinforcement. In nine tests by Ferreira1 and five by Birkle,10 EC2 predicts
Although there are only six results in Table 3, it is note- outside failures for slabs that actually failed in the shear-re-
worthy that, for all the methods of calculation, Vu/Vcalc inforced zones. Its predictions of failure modes in the other
decreases with increasing effective depth. This is not series are generally good. The main cause of the problem
surprising for ACI 318, which has no size factor, or for EC2, seems to be the overestimation of VR,cs. For the slabs by
where the size factor is constant for d ≤ 200 mm (7.9 in.). Ferreira,1 the mean Vu/VR,cs is 0.98, and the coefficient of
It is surprising for the CSCT, which includes a size factor variation is 0.061. For Birkle’s tests,10 the corresponding
taking account of the effective depth and the maximum size figures are 0.88 and 0.101, but would be improved if the
of aggregate. The best correlation in the table is that for four slabs with so/d less than 0.3 were excluded. The situ-
EC2*, which is the same as EC2, but without the limit on ation could be improved by either a reduction of VR,c or by
k = 1 + (200 / d ). With this modification, however, the interpreting the code’s expression for the design value of the
mean Vu/Vcalc is low, and the coefficient of 0.18 in Eq. (7) stud stress (fywd,ef) as not requiring a safety factor so long as
would need to be reduced. fywd,ef is less than fyw/1.15—that is, by taking fyw,ef as (250 +
Table 4 summarizes the results of the comparisons with 0.25d) ≤ fyw.
the 45 slabs with shear reinforcement, and all three methods The EC2 predictions of VR,out for slabs with radial arrange-
of calculation are broadly satisfactory. ments of shear reinforcement are generally satisfactory,
The coefficient of variation of Vu/Vcalc decreases with though perhaps over-conservative for the slabs by Gomes
increasing complexity in the method. The ACI method is and Regan.11 In these slabs, the 0.64d widths of the I-beam
the simplest and gives a coefficient of variation of 0.162. flanges reduced the clear tangential spacing of the shear
The EC2 method is slightly more complex, and reduces the reinforcement. This could be allowed for, and would make
coefficient by 0.026, while the CSCT is considerably more Vu/VR,out for these tests similar to those for other series.
complicated, but gives a further reduction of 0.015. The strength of Ferreira’s1 Slab C4 with an ACI cross
There are no unsafe predictions from ACI 318, but there arrangement of studs which failed inside is predicted very
are four from EC2 and the CSCT, with the lowest values conservatively with Vu/VR,out = 1.69. For Birkle’s10 slabs with
of Vu/Vcalc being 0.88 for EC2, and 0.90 for the CSCT. An the ACI layout, which failed outside, the strengths are well
overall reduction of Vcalc by 4% would make each of these predicted with Vu/VR,out = 1.21. Slab C4 was unrealistic in
methods safe in the sense of limiting the probability of an relation to EC2 design because it had six perimeters of studs,
unsafe prediction to 5%, assuming a statistically normal while the same strength would be calculated for a slab with
distribution of Vu/Vcalc. two perimeters of studs. The performance of C4 is in marked

ACI Structural Journal/March-April 2014 371


Fig. 14—Influence of slab rotation on Vu/Vcalc CSCT.
Fig. 15—Influence of slab rotation on shear strength.
contrast to that of slabs by Mokhtar et al.,12 with up to eight
perimeters of studs on stud rails. Their strengths are quite (refer to Carvalho et al.13). Reasonable results can, however,
well predicted by EC2. be obtained in most instances with fixed angles, provided the
expressions for VR,c and VR,s are constructed appropriately.
Influence of slab rotation on shear resistance This seems to be the case with ACI 318 and EC2, with the
provided by concrete former considering d/sr perimeters of studs acting at fyw, and
Unlike ACI 318 and EC2, the CSCT predicts that the shear the latter assuming 1.5d/sr perimeters acting at fyw,eff, which
resistance provided by concrete reduces with slab rotation, is typically approximately 0.7fyw for test slabs. The situation
which is assumed to be proportional to (V/Vflex)1.5. This has is more complex in the CSCT, as its numbers of perimeters
significant implications for design because Vu/Vflex may be depend on the exact distances of studs from the column.
close to one for practical slabs. Consequently, the CSCT can There are cases where the use of different failure surfaces
require significantly greater areas of shear reinforcement has a significant effect. Slabs 2, 3, 10, and 11 by Gomes
than EC2. The influence of slab rotation on Vu/Vcalc for the and Regan11 are an example. In these slabs so = sr ≅ 0.5d,
CSCT is illustrated by Fig. 14, where the ratio is plotted Asw = 226 mm2 (0.36 in.2) in Slabs 2 and 10, and 325 mm2
against the normalized rotation ψd/(16 + dg), with ψ calcu- (0.5 in.2) in Slabs 3 and 11. Slabs 2 and 3 had two perime-
lated by Eq. (13). There is a clear tendency for the CSCT ters of shear reinforcement, while Slabs 10 and 11 had three.
to become more conservative with increasing slab rotation, Thus, for ACI 318, two perimeters are taken into account for
which suggests that Vu is either independent of ψ, or that all the slabs, while in EC2, two perimeters are active in Slabs
the CSCT overestimates the influence of rotation. In the 2 and 3, but three are active in Slabs 10 and 11. All four
case of outside failure, this is to be expected as the rotation slabs failed inside their shear reinforced zones. The EC2
develops close to the column and not within a crack outside ratios Vu/VR,cs are 1.26 and 1.21 for Slabs 2 and 3, and 1.28
the shear reinforcement. and 1.31 for Slabs 10 and 11. The ACI ratios are 1.39 and
Muttoni4 plots Vu/u1d fc against ψd/(16 + dg) for 99 tests 1.28 for Slabs 2 and 3, and 1.58 and 1.61 for Slabs 10 and
of slabs without shear reinforcement, and shows that exper- 11, showing that the extra perimeter of shear reinforcement
imental strengths are close to the predictions of Eq. (15), had an effect.
which may appear to contradict the preceding paragraph. The same slabs illustrate a problem with the CSCT’s
considering the active shear reinforcement to be exactly
This, however, is not the case. Figure 15 shows Vu/u1d fc that within a distance d from a column rather than using an
plotted against ψd/(16 + dg) for slabs similar to Ferreira’s1 expression in d/sr. Because of variations of effective depths,
C2, but without shear reinforcement, and with ρ from 0.4 (so + sr) = 1.05d in three cases instead of the intended 1.0d.
to 4.0%. The values of Vu have been calculated by EC2 and This discrepancy has been ignored in the calculations for
the CSCT, u1 is the CSCT control perimeter, and ψ is the Table A1, but if the CSCT were applied strictly the ratios
rotation calculated by Eq. (13). It can be seen that the effect Vu/VR,cs, which are already unusually high for Slabs 10 and
shown in Muttoni’s figure can be accounted for by the EC2 11, would be significantly increased.
relationship between VR,c and ρ1/3 without involving ψ in the
calculations. CONCLUSIONS
Comparisons have been made between the punching
Failure surface and locations of shear strengths of 45 slabs with and six slabs without shear reinforce-
reinforcement ment and those predicted by ACI 318, EC2, and the CSCT.
ACI 318 and the CSCT assume punching surfaces to be ACI 318 is the simplest method, and gives only one unsafe
inclined at 45 degrees, while EC2 assumes an inclination of prediction, which is for a slab without shear reinforcement.
26.6 degrees. These are simplifications of a reality in which Its mean value of Vu/Vcalc for slabs with shear reinforcement
the angle increases with increasing shear reinforcement is rather high, and the coefficient of variation is 0.162. The

372 ACI Structural Journal/March-April 2014


most apparent weakness is the lack of any treatment of the c = side length of square column or diameter of circular column
d = mean effective depth
depth effect in the shear resistance from the concrete. dg = maximum size of aggregate
EC2 is only slightly more complicated, but reduces the dv = depth from tension reinforcement to compression zone
coefficient of variation of Vu/Vcalc by 0.026. The mean is anchorage of shear reinforcement
Es = modulus of elasticity of reinforcement
also reduced, and there are four unsafe predictions for slabs fc = cylinder compression strength of concrete
with and four for slabs without shear reinforcement. The fy = yield stress of flexural reinforcement
simplest way to obtain a characteristic level of safety would so = distance from column face to first perimeter of shear
reinforcement
be to reduce the constant in the expression for the concrete sr = radial spacing of shear reinforcement
component of resistance and extend the range of slab depths st = tangential spacing of shear reinforcement
affected by the depth factor. st,max = maximum value of st (general in outer perimeter of shear
reinforcement)
The CSCT is considerably more complex, and reduces u0 = length of column perimeter
the mean and coefficient of variation of Vu/Vcalc, by a further u1 = length of control perimeter for calculation of VR,c and VR,cs
0.015. There are unsafe predictions for four slabs with and uout = length of control perimeter for calculation of VR,out
uout,ef = effective value of uout for calculations by EC2, where st,max > 2d
one without shear reinforcement. V = applied shear force
The CSCT goes further than the other two approaches in Vcalc = calculated punching resistance
attempting to model the slab behavior. Although its expres- Vflex = flexural strength of slab calculated by yield-line theory
VR,c = punching resistance of slab without shear reinforcement
sion for total slab rotation seems good, the assumption that VR,cs = punching resistance within shear reinforced zone
half of this rotation occurs in the critical shear crack is incor- VR,max = maximum punching resistance for given column size, slab effec-
rect, as nearly all of it is at the column face. The assumption tive depth, and concrete strength
VR,out = punching resistance outside shear reinforced zone
that all critical cracks are at 45 degrees to the slab plane is VR,s = contribution of shear reinforcement to punching resistance VR,cs
also incorrect. Vu = experimental punching strength
The relationship assumed between the concrete compo- ρ = ratio of flexural reinforcement ρ = ρx ρy (calculated for width
nent of punching resistance and slab rotation is not of column plus 3d to either side in EC2)
ψ = rotation of part of slab outside critical shear crack
confirmed by the test data, and the determination of the area
of active shear reinforcement as that crossed by a particular
45 degree surface seems less satisfactory than considering REFERENCES
1. Ferreira, M. P., “Punção em Lajes Lisas de Concreto Armado com
d/sr perimeters. Armaduras de Cisalhamento e Momentos Desbalanceados,” PhD thesis,
Universidade de Brasília, Brasília, Brazil, 2010, 275 pp. (in Portuguese)
available at http://repositorio.bce.unb.br/handle/10482/8965.
AUTHOR BIOS 2. ACI Committee 318, “Building Code Requirements for Structural
ACI member Maurício P. Ferreira is a Lecturer at the Federal Univer-
Concrete (ACI 318-08) and Commentary,” American Concrete Institute,
sity of Para, Belem, Brazil. He received his PhD from the University of
Farmington Hills, MI, 2008, 473 pp.
Brasília, Brasília, Brazil, in 2010. His research interests include ultimate
3. Eurocode 2, “Design of Concrete Structures – Part 1-1: General Rules
shear design, strut and tie, and nonlinear finite element modeling.
and Rules for Buildings,” CEN, EN 1992-1-1, Brussels, Belgium, 2004,
225 pp.
ACI member Guilherme S. Melo is an Associate Professor at the Univer-
4. Muttoni, A., “Punching Shear Strength of Reinforced Concrete Slabs
sity of Brasilia, where he was Head of the Department of Civil and Environ-
without Transverse Reinforcement,” ACI Structural Journal, V. 105, No. 4,
mental Engineering. He is a member of ACI Committees 440, Fiber-Rein-
July-Aug. 2008, pp. 440-450.
forced Polymer Reinforcement; and Joint ACI-ASCE Committee 445, Shear
5. Fernadez-Ruiz, M., and Muttoni, A., “Applications of the Critical
and Torsion. His research interests include punching and post-punching of
Shear Crack Theory to Punching of R/C Slabs with Transverse Reinforce-
flat plates, the use of fiber-reinforced plastic (FRP) in concrete structures,
ment,” ACI Structural Journal, V. 106, No. 4, July-Aug. 2009, pp. 485-494.
and strengthening and rehabilitation of structures.
6. Fédération internationale du béton, “fib Model Code 2010, First
complete draft—V. 2,” Bulletin 56, fib, Lausanne, Switzerland, Apr. 2010,
ACI member Paul E. Regan is a Professor Emeritus at the University of
288 pp.
Westminster, London, UK, where he was Head of Architecture and Engi-
7. Regan, P. E., unpublished tests for RFA-TECH at Cambridge Univer-
neering. He was Chair of the European Concrete Committee (CEB) commis-
sity, 2009.
sion on member design. His research interests include member design in
8. Regan, P. E., and Samadian, F., “Shear Reinforcement against
both reinforced and prestressed concrete, with particular emphasis on
Punching in Reinforced Concrete Flat Slabs,” The Structural Engineer,
problems of punching, shear, and torsion.
V. 79, No. 10, May 2001, pp. 24-31.
9. Beutel, R., “Punching of Flat Slabs with Shear Reinforcement at Inner
ACI member Robert L. Vollum is a Reader in concrete structures at Impe-
Columns,” Rheinisch-Westfälischen Technischen Hochschule Aachen,
rial College London, London, UK, where he also received his MSc and PhD.
Aachen, Germany, 2002, 267 pp. (in German)
His research interests include design for shear, strut-and-tie modeling, and
10. Birkle, G., “Punching of Flat Slabs: The Influence of Slab Thickness
design for the serviceability limit states of deflection and cracking.
and Stud Layout,” PhD thesis, Department of Civil Engineering, University
of Calgary, Calgary, AB, Canada, Mar. 2004, 152 pp.
ACKNOWLEDGMENTS 11. Gomes, R., and Regan, P. E., “Punching Strength of Slabs Rein-
The authors are grateful to the Brazilian Research Funding Agencies forced for Shear with Offcuts of Rolled Steel I-Section Beams,” Magazine
CAPES (Higher Education Co-ordination Agency) and CNPq (National of Concrete Research, V. 51, No. 2, 1999, pp. 121-129.
Council for Scientific and Technological Development) for their support 12. Mokhtar, A. S.; Ghali, A.; and Dilger, W., “Stud Shear Reinforce-
throughout this research and to RFA-Tech for their permission to include ment for Flat Concrete Plates,” ACI Journal, V. 82, No. 5, Sept.-Oct. 1985,
test results from Reference 7. pp. 676-683.
13. Carvalho, A. L.; Melo, G. S.; Gomes, R. B.; and Regan, P. E.,
“Punching Shear in Post-Tensioned Flat Slabs with Stud Rail Shear
NOTATION Reinforcement,” ACI Structural Journal, V. 108, No. 5, Sept.-Oct. 2011,
Asw = area of shear reinforcement in one perimeter
pp. 523-531.

ACI Structural Journal/March-April 2014 373


APPENDIX A
Table A1—Comparison between theoretical and experimental results
Shear reinforcement Vu/Vcalc and critical strength
Slab Column d, fy, fc, fyw, s o, sr, stmax, Vu, Failure Vu/ CSCT
Author No. size, mm mm ρ, % MPa MPa Studs MPa mm mm mm kN mode Vflex ACI 318-08 EC2-04 average
C1 270 C 143 1.48 540 48 10 f10.0 x 6 535 70 100 436 858 In 0.72 1.34 Max 0.96 Out 1.07 In
C2 360 C 140 1.52 540 47 10 f10.0 x 6 535 70 100 464 956 In 0.78 1.27 Max 1.11 Out 1.12 In
C3 450 C 142 1.49 540 49 10 f10.0 x 6 535 70 100 491 1077 In 0.82 1.21 Out 1.20 Out 1.15 In
C4* 360 C 140 1.52 540 48 12 f10.0 x 6 535 70 100 900 1122 In 0.92 1.47 Max 1.69 Out 1.50 Out
C5 360 C 140 2.00 544 50 10 f10.0 x 6 535 70 100 464 1118 In 0.88 1.44 Max 1.16 Out 1.29 In
Ferreira1

C6 360 C 143 1.48 540 49 10 f10.0 x 6 535 70 100 464 1078 In 0.86 1.36 Max 1.19 Out 1.24 In
C7 360 C 144 1.47 540 49 10 f10.0 x 7 535 55 80 442 1110 In 0.88 1.39 Max 1.21 Out 1.09 Out
C8 360 C 144 1.47 540 48 12 f10.0 x 6 535 70 100 388 1059 In 0.84 1.34 Max 1.03 Out 1.14 In
S1 300 S 145 1.46 540 48 12 f10.0 x 2 535 70 100 177 1022 Out 0.80 1.71 In 1.36 Out 1.37 Out
S2 300 S 143 1.48 540 49 12 f10.0 x 4 535 70 100 280 1128 In 0.89 1.77 Out 1.12 Out 1.24 Out
S5 300 S 143 1.48 540 50 — — — — — 779 P 0.61 1.30 P 1.20 P 1.24 P
S7 300 S 143 1.48 540 49 12 f12.5 x 4 518 70 100 280 1197 Out 0.94 1.88 Out 1.19 Out 1.32 Out
1 300 S 150 1.45 550 33 10 f10.0 x 4 550 80 120 390 881 — 0.79 1.45 Out 1.02 Out 1.06 In
2 300 S 150 1.76 550 30 12 f10.0 x 6 550 60 100 390 1141 — 0.88 1.71 Out 1.13 Out 1.11 Out
Regan7

3 300 S 150 1.76 550 26 10 f12.0 x 5 550 60 120 455 1038 — 0.83 1.73 Out 1.22 Out 1.09 Out
5 240 C 160 1.65 550 62 12 f12.0 x 5 550 80 120 352 1268 — 0.88 1.61 Max 0.88 Out 1.12 In
6 240 C 150 1.75 550 42 12 f10.0 x 5 550 75 120 349 1074 — 0.83 1.81 Max 1.04 In 1.24 In
R3 200 S 160 1.26 670 33 8 f12.0 x 4 442 80 120 413 850 Out 0.63 1.44 Out 1.04 Out 0.90 Out
Samadian8
Regan and

R4 200 S 160 1.26 670 39 8 f12.0 x 6 442 80 80 444 950 Out 0.69 1.39 Out 1.10 Out 0.93 Out
A1 200 S 160 1.64 570 37 8 f10.0 x 6 519 80 80 444 1000 Out 0.67 1.50 Out 1.08 Out 0.98 Out
A2 200 S 160 1.64 570 43 8 f10.0 x 4 519 80 120 413 950 In 0.62 1.42 Out 1.03 In 1.00 In
Z1 200 C 250 0.80 890 25 12 f14.0 x 5 580 100 200 518 1323 Max 0.41 1.50 Max 1.26 Max 0.96 In
Z2 200 C 250 0.80 890 26 12 f14.0 x 5 580 88 200 511 1442 Max 0.44 1.59 Max 1.30 Max 1.08 In
Beutel9

Z3 200 C 250 0.80 890 24 12 f14.0 x 5 580 95 188 487 1616 Max 0.50 1.86 Max 1.57 Max 1.20 In
Z4 200 C 250 0.80 890 32 12 f14.0 x 5 580 88 175 459 1646 Max 0.49 1.66 Max 1.27 Max 1.18 In
Z5 263 C 250 1.25 562 28 12 f16.0 x 5 544 94 188 505 2024 Max 0.41 1.90 Max 1.31 Max 1.28 In
Z6 200 C 250 1.25 562 37 12 f16.0 x 5 544 94 188 489 1954 Max 0.39 1.81 Max 1.31 Max 1.23 In
1 200 S 159 1.27 680 40 — — — — — 560 P 0.40 1.16 P 0.94 P 1.00 P
1a 200 S 159 1.27 680 41 — — — — — 587 P 0.41 1.20 P 0.98 P 1.03 P
2* 200 S 153 1.32 680 34 8 f6.0 x 2 430 80 80 255 693 In 0.53 1.64 In 1.26 In 1.12 Out
3* 200 S 158 1.27 670 39 8 f6.9 x 2 430 80 80 255 773 In 0.57 1.64 In 1.21 In 1.20 Out
Gomes and Regan11

4* 200 S 159 1.27 670 32 8 f8.0 x 3 430 80 80 368 853 Out 0.64 1.98 In 1.27 Out 1.26 Out
5* 200 S 159 1.27 670 35 8 f10.0 x 4 430 80 80 481 853 Out 0.63 1.77 Out 1.24 Out 1.13 Out
6 200 S 159 1.27 670 37 8 f10.0 x 4 430 80 80 323 1040 Out 0.76 2.07 Out 1.23 Out 1.34 Out
7 200 S 159 1.27 670 34 8 f12.0 x 5 430 80 80 385 1120 Out 0.83 2.02 Out 1.38 Out 1.38 Out
8 200 S 159 1.27 670 34 8 f12.0 x 6 430 80 80 447 1200 Out 0.89 1.90 Out 1.48 Out 1.38 Out
9 200 S 159 1.27 670 40 8 f12.2 x 9 430 80 80 425 1227 — 0.89 1.31 Out 1.09 Out 1.26 Max
10 200 S 154 1.31 670 35 8 f6.0 x 5 430 80 80 385 800 In 0.61 1.58 In 1.28 In 1.33 In
11 200 S 154 1.31 670 35 8 f6.9 x 5 430 80 80 385 907 In 0.70 1.68 Out 1.31 In 1.42 In
1 250 S 124 1.53 488 36 — — — — — 483 P 0.56 1.30 P 1.11 P 1.10 P
2* 250 S 124 1.53 488 29 8 f9.5 x 6 393 45 90 721 574 In 0.68 1.24 Out 1.19 Out 1.08 Out
3 250 S 124 1.53 488 32 8 φ9.5 x 6 393 45 90 495 572 In 0.67 1.10 Out 1.12 Out 1.02 In
4* 250 S 124 1.53 488 38 8 φ9.5 x 5 465 30 60 403 636 Out 0.73 1.67 In 1.21 Out 1.09 Out
5* 250 S 124 1.53 488 36 8 φ9.5 x 5 465 30 60 403 624 Out 0.72 1.67 In 1.21 Out 1.09 Out
Birkle10

6 250 S 124 1.53 488 33 8 φ9.5 x 5 465 30 60 330 615 Out 0.72 1.67 Out 1.18 Out 1.04 Out
7 300 S 190 1.29 531 35 — — — — — 825 P 0.49 1.12 P 0.94 P 1.02 P
8* 300 S 190 1.29 531 35 8 φ9.5 x 5 460 50 100 658 1050 In 0.62 1.29 Out 0.98 Out 0.97 In
9* 300 S 190 1.29 531 35 8 φ9.5 x 6 460 75 150 1188 1091 In 0.64 1.28 In 1.06 In 1.15 In
10* 350 S 260 1.10 524 31 — — — — — 1046 P 0.40 0.88 P 0.78 P 0.86 P
11* 350 S 260 1.10 524 30 8 φ12.7 x 5 409 65 130 856 1620 In 0.63 1.24 Out 1.00 Out 1.02 In
12* 350 S 260 1.10 524 34 8 φ12.7 x 6 409 95 195 1541 1520 In 0.58 1.03 In 0.90 Out 1.08 In
*
ACI stud layout.
Notes: Vu includes self-weight; Vflex is approximate yield-line capacity from Eq. (14).
Shear reinforcement: In References 1 and 7: deformed studs, 3φ heads, so as given for all lines; in Reference 8, slabs R, plain studs, 2.5φ heads, Slabs A deformed studs, 2.5φ
heads, so as given for orthogonal lines, so = 40 mm for diagonal lines; in Reference 9, deformed studs, 3φ heads, so as given for all lines; in Reference 11, I-beam slices, flange
breath 102 mm, web breath 4.7 mm. φ values in the table are equivalent diameters giving the same areas as the actual web sections. so as given for orthogonal lines, so = 40 mm for
diagonal lines; in Reference 10, plain studs with 3.2φ heads, so as given for all lines. Birkle’s Slabs 5 and 6 had 7 perimeters of studs. The outer two, with sr = d, have been ignored.
Aggregate (maximum size and type): In Reference 1, 9.5 mm crushed limestone. In References, 7, 8, 9, and 11, 20 mm gravel. In Reference 10, Slabs 1-6—14 mm; Slabs
7-12—20 mm, type unknown.
Failure modes: P is punching of slabs without shear reinforcement, In = failure inside shear reinforced zone (VR,cs), Out = failure outside shear reinforced zone (VR,out); Max =
inclined compression failure of concrete close to column (VR,max); in Reference 7 and Slab 9 of Reference 10, the concrete soffit around the column crushed and spalled due to
tangential compression, the spalling extended and at failure there was inclined cracking starting at the end of the spalled area.
1 mm = 0.03937 in.; 1 kN = 0.225 kip; 1 MPa = 145 psi.

374 ACI Structural Journal/March-April 2014


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 111-S33

Behavior of Concentrically Loaded Fiber-Reinforced


Polymer Reinforced Concrete Columns with Varying
Reinforcement Types and Ratios
by Hany Tobbi, Ahmed Sabry Farghaly, and Brahim Benmokrane
Fiber-reinforced polymer (FRP) materials have proven their effec- Columns are one of several structural elements that may
tiveness as an alternative reinforcement for concrete structures in be exposed to severe environmental effects. The response of
severe environmental conditions. Many studies have investigated FRP bars in compression is affected by different modes of
the flexural and shear behaviors of FRP-reinforced concrete beams failure (transverse tensile failure, buckled FRP bar, or shear
and slabs. Limited research, however, has gone into investigating
failure). General acceptance of FRP bars by practitioners
the behavior of internally reinforced FRP concrete columns. This
requires that appropriate design guidelines for using FRP bars
paper reports the experimental investigation of the compressive
performance of concrete columns reinforced longitudinally with in compression members be established. Due to the lack of
FRP or steel bars and with FRP as transverse reinforcement. experimental data, the current ACI 440.1R (ACI Committee
Twenty concrete columns measuring 350 x 350 x 1400 mm (13.8 x 440 2006) design guidelines do not recommend the use of
13.8 x 55.1 in.) were constructed and tested under concentric FRP bars as longitudinal reinforcement in compression
compressive load. The parametric study included variables such as members, while the CSA S806 (2012) code states that the
transverse reinforcement configuration, material type and spacing, compressive contribution of FRP longitudinal reinforcement
longitudinal reinforcement ratio, and confining volumetric stiff- is negligible. Moreover, confined concrete behaves differ-
ness. Results showed that FRP bars have contribution as longitu- ently from unconfined concrete. Several studies clarified
dinal reinforcement for concrete columns subjected to concentric the importance of the lateral reinforcement as a confining
compression and that the combination of FRP transverse rein-
system to the performance in terms of capacity and ultimate
forcement and steel longitudinal bars offers acceptable strength
axial strain of axially loaded concrete members (Richart et
and ductility behavior.
al. 1928, 1929; Sheikh and Uzumeri 1980, 1982; Sheikh
Keywords: column; concentric compression; confinement volumetric 1982; Saatcioglu and Razvi 1992; Mander et al. 1988 a,b;
stiffness; failure mechanism; fiber-reinforced polymer (FRP); steel; Teng et al. 2002; Harries and Kharel 2003).
volumetric ratio.
RESEARCH SIGNIFICANCE
INTRODUCTION As the use of FRP reinforcement in concrete structures
The deterioration of infrastructure owing to corrosion of grows, appropriate design guidelines for axially concen-
steel reinforcement is one of the major challenges facing the tric loaded concrete columns should be established. In
construction industry. The use of reinforcement with fiber- this regard, laboratory investigations should be conducted
reinforced polymer (FRP) composite materials in concrete to expand understanding of the compressive behavior of
structures subjected to severe environmental exposure has concrete columns internally reinforced with FRP, particu-
been growing to overcome the common problems caused by larly given the lack of data about this application, but also to
corrosion of steel reinforcement (ACI Committee 440 2007; highlight the most important parameters affecting compres-
Federation Internationale de Béton 2007). Recent advances sive performance of FRP reinforced columns. This study
in polymer technology have led to the development of the investigated concrete columns reinforced longitudinally
latest generation FRP reinforcing bars (ACI Committee 440 with glass FRP (GFRP), carbon FRP (CFRP), and steel bars
2007). These corrosion-resistant bars have shown promise and GFRP and CFRP transverse reinforcement subjected to
as a way to further protect bridges and public infrastructure concentric loading. The experimental study yielded a better
from the devastating effects of corrosion. With standards understanding of the mechanical behavior of FRP reinforced
ACI 440.6M (ACI Committee 440 2008) and CSA S807 columns. The set of specimens is presented to enrich the
(2010) and bars being produced of the highest quality, FRP literature regarding the use of FRP as internal reinforcement
bars are emerging as a realistic and cost-effective reinforce- for compressive members in preparation for developing
ment alternative to traditional steel for concrete structures design models.
under severe environmental conditions. Steel bars cannot,
however, simply be replaced with FRP bars due to various
differences in the mechanical and bond properties of the
two materials (Nanni 1993; ISIS Canada Research Network ACI Structural Journal, V. 111, No. 2, March-April 2014.
2007) and the greater variation of material properties for MS No. S-2012-123.R1, doi:10.14359.51686528, was received November 9, 2012,
and reviewed under Institute publication policies. Copyright © 2014, American
FRP reinforcing products. Concrete Institute. All rights reserved, including the making of copies unless
permission is obtained from the copyright proprietors. Pertinent discussion including
author’s closure, if any, will be published ten months from this journal’s date if the
discussion is received within four months of the paper’s print publication.

ACI Structural Journal/March-April 2014 375


Fig. 1—(a) and (b) C-shaped; and (c) closed transverse reinforcement.

Fig. 2—Transverse reinforcement configuration: (a) 1; (b) 2; and (c) 3.


EXPERIMENTAL PROGRAM were used: the first is assembled from C-shaped parts as
The experimental study comprised 20 concrete columns shown in Fig. 1(a) and (b), and the second is closed form,
measuring 350 x 350 x 1400 mm (13.8 x 13.8 x 55.1 in.) as shown in Fig. 1(c). The first shape was the first avail-
subjected to concentric compressive loading. These dimen- able product for bent bars; therefore, it has been chosen to
sions are representative of the columns commonly found build up the transverse reinforcement of the FRP reinforced
in concrete structures. One column was kept un-reinforced columns. The second shape, however, has the benefit of
(plain concrete) while the remaining 19 were internally rein- eliminating the discontinuity of the C-shaped and reducing
forced with FRP and steel according to different parameters. the assembly labor time. Generally speaking, both can be
All used transverse reinforcements were GFRP or CFRP, used in construction. The closed transverse reinforcements
while the longitudinal reinforcement was GFRP, CFRP, or were cut from continuous square spirals, with an overlap
steel bars. equal to one side length. For each transverse reinforcement
Studied parameters included the shape of transverse rein- shape, three configurations labeled 1, 2, and 3, as shown
forcement (C-shaped parts assembly or closed ties, as shown in Fig. 2 (a), (b), and (c), were investigated in Tobbi et al.
in Fig. 1), longitudinal reinforcement ratio, longitudinal (2012), which revealed that the effect of Configuration 2 is
reinforcement material (GFRP, CFRP, or steel), FRP-trans- not different from Configuration 3; therefore, only Configu-
verse reinforcement material (GFRP or CFRP), the diameter rations 1 and 3 will be addressed in this study. In the case of
of the transverse reinforcement (No. 9.5 and 12.7 mm [No. 3 C-shaped transverse reinforcements in Configuration 1, the
and No. 4]), transverse reinforcement spacing, and confining cross hairpins in two consecutive layers were staggered to
volumetric stiffness. Volumetric ratio is an important param- preserve overall symmetry. In the case of the closed trans-
eter for confinement efficiency (Sheikh and Uzumeri 1980, verse reinforcements in Configuration 1, cross hairpins were
1982; Sheikh 1982; Saatcioglu and Razvi 1992; Mander et staggered C-shaped legs. In Configuration 3, the cross ties
al. 1988 a,b; Watson et al. 1994; Cusson and Paultre 1994; were closed rectangular ties with one side overlap. All CFRP
Saatcioglu et al. 1995) for passive confinement with internal transverse reinforcements were closed with different diame-
reinforcement. Volumetric ratio ρv is defined as the ratio ters (No. 9.5 and 12.7 mm [No. 3 and No. 4]) to investigate
of the volume of transverse confining reinforcement to the their effect on the compressive performance of the columns.
volume of confined concrete core. FRP mechanical propri- The test matrix is listed in Table 1.
eties vary depending on fiber material and fiber content. For To fulfill the objectives of the parametric study, eight
consistency, therefore, volumetric ratio should be multi- columns were exclusively reinforced with FRP: seven with
plied by the modulus of elasticity of FRP confining material GFRP, both longitudinal bars and transverse reinforcements
(ρv × Ef), which is the so-called confining volumetric stiff- (Configurations 1 and 3), and one with CFRP, also with both
ness. The same confining volumetric stiffness (ρv × Ef) can longitudinal bars and transverse reinforcements. Eleven
be obtained by changing at least two of the following param- columns were reinforced longitudinally with steel bars and
eters: transverse reinforcement’s configuration, spacing, FRP transverse reinforcements (Configurations 1 and 3):
material, or diameter. four reinforced transversally with GFRP, and seven with
The GFRP transverse reinforcement diameter was CFRP. One column was unreinforced (plain concrete).
12.7 mm (No. 4). Two transverse reinforcement shapes

376 ACI Structural Journal/March-April 2014


Table 1—Test matrix
Longitudinal reinforcement Transverse reinforcement Ties
fc′, ΜPa Ties spacing, ρv x Ef,
Group Specimen (ksi) Designation Material Designation Material configuration mm (in.) GPa
P-0-00-0 — — — — — — —
G-1c-120-1.9 8 No. 6 (19 mm) 1 120 (4.7) 0.96
Columns entirely reinforced with

33 (4.8)
G-3c-120-1.9 12 No. 5 (15.9 mm) 3 120 (4.7) 1.28
G-3c-80-1.9 12 No. 5 (15.9 mm) 3 80 (3.2) 1.92
G-1-120-1.9 8 No. 6 (19 mm) 1 120 (4.7) 1.18
FRP bars

GFRP No. 4 (12.7 mm) GFRP


G-3-120-1.9 12 No. 5 (15.9 mm) 3 120 (4.7) 1.58
4 No. 4 + 4 No. 5
G-1-120-1.0 1 120 (4.7) 1.18
(12.7; 15.9 mm)
G-1-120-0.8 35 (5.1) 8 No. 4 (12.7 mm) 1 120 (4.7) 1.18
No. 3C-1-67-1.6 2 x 8 No. 4 (12.7 mm)* CFRP No. 3 (9.5 mm) CFRP 1 67 (2.6) 2.87
G-1-120-1.0S 4 M15 + 4 M10† GFRP 1 120 (4.7) 1.18
Columns reinforced with steel longitudinal bars

No. 4 (12.7 mm)


C-1-120-1.0S 4 M15 + 4 M10 CFRP 1 120 (4.7) 3.04
No. 3C-1-67-1.0S 4 M15 + 4 M10 1 67 (2.6) 2.87
and FRP tie and cross-ties

No. 3 (9.5 mm) CFRP


No. 3C-3-80-1.0S 12 M10 3 80 (3.2) 3.20
C-1-80-1.0S 4 M15 + 4 M10 1 80 (3.2) 4.56
C-1-60-1.0S 4 M15 + 4 M10 Steel 1 60 (2.4) 6.08
CFRP
C-3-120-1.0S 12 M10 3 120 (3.2) 4.05
27 (3.9)
C-3-80-1.0S 12 M10 No. 4 (12.7 mm) 3 80 (3.2) 6.08
G-1-80-1.0S 4 M15 + 4 M10 1 80 (3.2) 1.78
G-3-120-1.0S 12 M10 GFRP 3 120 (4.7) 1.58
G-3-80-1.0S 12 M10 3 80 (3.2) 2.37
*
Bundled bars.

M15 placed at corners.
Notes: P is plain concrete; G is GFRP; C is CFRP; (1; 3) = Stirrup configuration; c is «C» shaped legs assembly; (120; 80; 67; 60) is stirrup spacing, mm; (0.8; 1.0; 1.9) is longitu-
dinal reinforcement ratio; S is steel longitudinal bars.

Materials
Table 2—Tensile properties of FRP and steel
The columns were cast vertically using normalweight
longitudinal reinforcement
ready mixed concrete with a target 28-day concrete compres-
sive strength of 30 MPa (4.4 ksi). The columns were cured Af, mm2 Ef, GPa ffu, MPa
for 7 days, after which the specimens were left in the labo- Bar type db, mm (in.) (in.2) (ksi) (ksi) εf, %
ratory at ambient temperature for at least three more weeks No. 4 127 46.3 1040
12.7 (0.5) 2.25
before testing. The concrete compressive strength used for GFRP (0.19) (6715) (151)
analysis was based on the average values of tests performed No. 5 199 48.2 751
on at least five 150 x 300 mm (6 x 12 in.) cylinders for each 15.9 (0.62) 1.56
GFRP (0.31) (6990) (109)
concrete batch under displacement control standard rate of
No. 6 284 47.6 728
0.01 mm/s (3.9 × 10–4 in./s) (Table 1). Grade 60 steel rein- GFRP
19.1 (0.75)
(0.44) (6904) (106)
1.53
forcing bars were used as longitudinal reinforcement for
No. 4 127 137 1902
specific specimens. Table 2 provided the tensile properties CFRP
12.7 (0.5)
(0.19) (19,870) (276)
1.38
of Grade 60 steel bars.
The longitudinal reinforcement for the exclusively FRP- 100 200 fy = 450
Steel M10 11.3 (0.44) εy = 0.2
(0.15) (29,000) (65)
reinforced columns was (1) No. 12.7 mm (No. 4) straight
CFRP bars, and (2) No. 15.9 mm (No. 5) and 19.1 mm (No. 6) Steel M15 16.0 (0.62)
200 200 fy = 460
εy = 0.2
GFRP straight bars. The tensile properties of longitudinal FRP (0.31) (29,000) (66)
and steel bars were determined by performing the B.2 test Notes: db is bar diameter; Af is cross-sectional area of bar; Ef is modulus of elasticity
method according to ACI 440.3R (ACI Committee 440 2004) of bar; ffu is ultimate tensile strength of bar; εf is ultimate strain of bar.
as reported in Table 2. Bent bars of 12.7 mm (No. 4) GFRP
to the ACI 440.3R B.2 test method (ACI Committee 440
and 9.5 mm (No. 3) and 12.7 mm (No. 4) CFRP were used
2004). The ultimate bent strength ffu,bend, however, was deter-
as transverse reinforcements. The ultimate tensile strength
mined using the B.5 test method according to ACI 440.3R
ffu and modulus of elasticity Ef for the straight portions of
(ACI Committee 440 2004). Table 3 provides the measured
the transverse reinforcements were determined according

ACI Structural Journal/March-April 2014 377


Fig. 3—(a) Loading machine; and (b) instrumentation.

Table 3—Tensile properties of FRP transverse


reinforcement
Straight portion Bend portion
ffu, MPa ffu,bend, ffu,bend/
Bar type (ksi) Ef, GPa (ksi) εf, % MPa (ksi) ffu
No. 3 CFRP 1327 (192) 126 (18,275) 1.05 614 (89) 0.46
No. 4 CFRP 1372 (198) 133 (19,290) 1.03 700 (101) 0.51
No. 4 GFRP 962 (139) 52 (7542) 1.85 500 (72) 0.52
C-shaped
640 (92) 44 (6382) 1.45 400 (58) 0.62
No. 4 GFRP

Notes: Ef is modulus of elasticity of bar; ffu is ultimate tensile strength of bar; εf is


ultimate strain of bar; ffu,bend is ultimate tensile strength of bar bend.

column specimens was recorded using four linear variable


Fig. 4—Stress-strain relationship for both plain concrete differential transducers (LVDTs) located at the midheight
cylinder and P-0-00-0 column. (Note: 1 MPa = 0.145 ksi.) of each side of the specimens, as shown in Fig. 3(b). The
top and bottom ends of the specimens were capped with a
tensile strength and modulus of elasticity for the straight and
thin layer of high-strength mortar to ensure that the bearing
bent portions. The GFRP and CFRP longitudinal bars were
surfaces were parallel and the load was distributed uniformly
pultruded products (Pultrall, Inc. 2009). Transverse rein-
during testing. To ensure the failure would occur in the
forcements were fabricated with the bend process (Pultrall,
instrumented region, the ends of each specimen were further
Inc. 2009). Sand coating was used on the surface of the
confined with bolted steel plates made from 13 mm (0.5 in.)
longitudinal and transverse FRP bars to improve the bond to
thick steel plates (Fig. 3(b)).
concrete, as in standard industry practice.
EXPERIMENTAL RESULTS AND DISCUSSION
Instrumentation and testing procedures
Reinforcement strain was measured with electrical strain
Overall behavior
gauges adhered to the bars at midheight of the column. A
The unreinforced plain concrete column (P-0-00-0) was
set of ties in each specimen was instrumented with strain
the first to be tested. The stress-strain behavior until peak
gauges placed at the middle and in the corner of the outer tie
was similar to the concrete cylinder, while peak stress was
and the cross hairpins. The test specimens were loaded by a
slightly lower as shown in Fig. 4. Post-peak behavior was
rigid MTS high force load frame (Fig. 3(a)) with a maximum
completely different. Concrete cylinders exhibit a consid-
compressive capacity of 11,400 kN (2,560,000 lbf) having
erable softening branch meaning gradual damage, while
the load controlled up to 2200 kN (495,000 lbf) with a rate
P-0-00-0 failure was brittle; total loss of sustained load
of 2.5 kN/s (562 lb/s). Thereafter, displacement control
occurred just after reaching peak stress. This difference in
was used to apply the load until failure with the rate of
post-peak behavior was due to the higher energy accumu-
0.002 mm/s (7.87 × 10–5 in./s). The axial displacement of the

378 ACI Structural Journal/March-April 2014


Fig. 5—Total cross-section-based stress-strain curves for all tested columns.
lated by the P-0-00-0 column compared to the small concrete of transverse reinforcing associated with complete spalling
cylinder. The post-peak behavior of plain concrete subjected of concrete cover, and ended with total failure of the column
to axial load is an important parameter for determination of (Fig. 6(c)). This phase is clearly governed by transverse
cover contribution of the reinforced columns subjected to reinforcement and longitudinal bar material (Fig. 5). The
the same type of load. The stress-strain curve for confined following sections contain detailed discussion.
concrete core of all reinforced columns is then obtained.
Considering the total cross-sectional area for concrete Confined concrete core behavior
columns with FRP transverse reinforcements (Fig. 5), the Analyzing compressive behavior of internally confined
stress-strain curve was divided into three phases. The first concrete columns considering the total cross-sectional area
phase corresponds to the behavior until the peak stress, from the starting of the elastic phase until failure is not
which was similar to plain concrete column, implying that accurate because the stress calculation does not take into
transverse reinforcement had no effect on this phase. The account the degradation of concrete cover contribution
concrete cover was visually free of cracks (Fig. 6(a)), yet after cracking. Nevertheless, when concrete cover is spalled
the peak stress varied depending on the longitudinal rein- off, the confined concrete core remained uncracked until a
forcement material and ratio (Fig. 5(b) and (c)). The second certain level, depending on the confinement effect. There-
phase was very short, and was characterized by a rapid drop fore, studying the effect of confinement accurately requires
in bearing capacity. This started once the peak stress was considering the confined concrete core only. Nevertheless,
reached, and finished with passive-confinement activation the concrete cover strength should be subtracted from the
(strain increase in the transverse reinforcements). In this total applied load based on the behavior of the plain concrete
phase, cracks began growing in the concrete cover, as shown column, as shown in Fig. 4. The reduced load, divided by the
in Fig. 6(b), leading to gradual spalling that reduced the concrete core area delimited by the centerlines of the outer
load-resisting cross-sectional area and resulted in strength transverse reinforcements, presents the column’s actual
degradation. The third phase is characterized by activation stress behavior. Figure 7 shows the curves representing the

ACI Structural Journal/March-April 2014 379


Fig. 6—Cracking appearance of test specimens at different loading stages.
Strength and failure mode
Different failure modes were observed based on rein-
forcement layout. The failure mode of exclusively FRP-
reinforced column followed this progression: 1) crushing or
buckling of the FRP longitudinal bar; and 2) transverse rein-
forcement rupture. Excessive buckling of the longitudinal
steel bars was the failure mode of columns reinforced longi-
tudinally with steel bars. The failure modes were governed
by the shape, configuration, and diameter of the transverse
reinforcements, as well as longitudinal bar material.
The failure of all longitudinally and transversally FRP-
reinforced columns was due to longitudinal bar crushing or
buckling, as shown in Fig. 8. In general, the columns with
C-shaped GFRP transverse reinforcements experienced
brittle failure. The failure of Column G-1c-120-1.9, which
had the lowest confinement volumetric stiffness, was explo-
sive. Column G-3c-120-1.9, which had higher confinement
Fig. 7—Effect of concrete cover (G-3c-80-1.9). volumetric stiffness, showed failure starting with the longi-
tudinal GFRP bars crushing, followed by total concrete
axial stress sustained by the concrete with respect to: 1) the crushing. In both columns, the slipping of the outer C-shaped
total load divided by the total concrete area (Path 0-A-B′-C′); transverse reinforcements at the splice location occurred due
and 2) the total load divided by the confined concrete area to concrete core expansion pressure, leading to degradation
delineated by the centerline of the outer transverse reinforce- of sustained load until crushing of the longitudinal bars,
ments (Path 0-A′-B-C). The actual response of the concrete followed instantaneously by concrete core crushing. More-
column, represented by the bold curve (Path 0-A-B-C), is over, inclined shear sliding surfaces were observed, leading
expected to be a combination of the two calculated curves. to a separation of the concrete core into two wedges, causing
The response of the concrete column (bold curve) coincides a sudden drop in axial strength. The failure of Column
with the ascending part of the lower curve (total concrete G-3c-80-1.9 was different: no transverse reinforcement slip-
area) up to Point A, which corresponds to the spalling of page was observed. This might be attributed to the smaller tie
the concrete cover. When the concrete cover no longer spacing, which allowed the column to fail progressively in
contributed to the axial strength, the response of the concrete successive crushing of all longitudinal GFRP bars followed
column coincided with the part of the higher curve (confined by concrete core crushing. The FRP reinforced columns with
concrete area) that follows Point B, when the concrete core closed transverse reinforcements failed progressively due to
began to gain strength due to confinement by the trans- successive crushing of the longitudinal bars before concrete
verse reinforcement. The transition between Points A and core crushing. No transverse reinforcements rupture was
B of the response of the concrete column was quantified by observed, except in Column G-3-120-1.9, which was the
subtracting the contribution of the concrete cover (which only FRP-reinforced column with Configuration 3 trans-
decreased with increasing axial deformation) based on the verse reinforcements.
stress-strain response of the plain concrete (Fig. 4). Point C Reinforcing the columns longitudinally with steel bars
corresponds to the ultimate strength of the tested columns. instead of FRP bars changed the failure mode. The longi-

380 ACI Structural Journal/March-April 2014


Fig. 8—Failure mode of columns reinforced longitudinally and transversely with FRP.

Fig. 9—Failure mode of columns reinforced longitudinally with steel and transversely with FRP.
tudinal steel bars consistently buckled (Fig. 9(a) and (b)).
In addition, cross-tie rupture was observed for columns
with Configuration 3, (Fig. 9(c)), while failure was due to
excessive bars buckling and substantial decrease in bearing
capacity in columns with Configuration 1 transverse rein-
forcements. Moreover, excessive buckling of the longitu-
dinal bars in columns induced openings in the GFRP trans-
verse reinforcements, as shown in Fig. 9(b). Opening (albeit
minor) was also observed with transverse CFRP reinforce-
ment at 80 and 60 mm (3.2 and 2.4 in.) spacing.
Failure due to transverse reinforcement rupture was
experienced in columns transversely reinforced with CFRP
No. 9.5 mm (No. 3) with both Configurations 1 and 3
(Fig. 10). Even Column No. 3C-1-67-1.6 experienced trans- Fig. 10—Transverse reinforcement rupture for No. 3 CFRP
verse reinforcement rupture after the longitudinal CFRP laterally reinforced columns.
bars experienced crushing.
strength fc′. Columns were cast in three different groups.
Parametric investigation While the targeted concrete strength was 30 MPa (4.4 ksi),
Parametric investigation was carried out to study the the actual concrete strength for the three groups was 33,
strength mechanism and performance based on stress-strain 35, and 27 MPa (4.8, 5.1, and 3.9 ksi). Before testing the
relationship for the tested columns. The investigated param- columns, actual cross-sectional area was measured to calcu-
eters included transverse reinforcement shape, material, late the precise stress values.
spacing and diameter (No. 9.5 and 12.7 mm [No. 3 and
No. 4]), longitudinal reinforcement ratio, longitudinal rein- Effect of transverse reinforcement shape
forcement material and confining volumetric stiffness. (C-shaped versus closed)
To compare the strength behavior of columns cast from Four columns were studied to investigate the effect
different concrete batches, the stress values σc were normal- of transverse reinforcement shape. Two columns, G-1c-
ized to the cylinder compressive strength fc′ of the batch. 120-1.9 and G-3c-120-1.9, were transversely reinforced
Therefore, the stress response σc along the test for each with C-shaped GFRP No.12.7 mm (No. 4). The other two
column was divided by the concrete compressive cylinder columns, G-1-120-1.9 and G-3-120-1.9, were transversely

ACI Structural Journal/March-April 2014 381


Fig. 12—Effect of longitudinal reinforcement on compres-
Fig. 11—C-shaped Vs closed transverse reinforcement
sive behavior of columns.
normalized stress-strain relationship.
(1.0%). Two main differences were observed: the peak stress
reinforced with closed GFRP No.12.7 mm (No. 4). The four
for the steel longitudinally reinforced column was higher than
columns had identical longitudinal reinforcement.
that of GFRP-reinforced column, and the GFRP-reinforced
Figure 11 shows the normalized stress-strain response.
column showed stabilization of the load-carrying capacity,
The confined concrete strength gain (fcc′/fc′) is quite similar
represented by nearly horizontal plateau until failure, at the
for the columns transversely reinforced with either trans-
post-peak phase, while the load-carrying capacity of steel
verse reinforcement type. Moreover, Configuration 3
longitudinally reinforced column decreased after reaching
showed a higher strength gain than Configuration 1. Never-
the peak load. The difference in post-peak behavior is due to
theless, a significant difference was noted based on trans-
longitudinal reinforcement material. Indeed, the load carried
verse reinforcement type. In the columns with C-shaped
by steel bars after yielding remained constant, while the load
transverse reinforcement (G-1c-120-1.9 and G-3c-120-1.9),
increased with axial strain with the elastic GFRP bars. This
the strength decreased due to leg slippage after reaching
behavior is more pronounced in Column G-3c-80-1.9, which
normalized confined concrete strength (fcc′/fc′). Therefore,
was the only column that exhibited stiffening behavior after
the normalized ultimate strength (fcu′/fc′) corresponding to
cracking of the cover and a reduction in the load drop that
ultimate axial strain εcu was lower than the normalized peak
ended with a second peak load higher than the first one, as
strength (fcc′/fc′). Meanwhile, in the columns with closed
shown in Fig. 5(a).
transverse reinforcement (G-1-120-1.9 and G-3-120-1.9),
The ultimate axial strain of the columns reinforced longi-
no descending branch was observed. The failure of columns
tudinally with FRP is lower than those reinforced with steel.
occurred at the maximum normalized confined concrete
Columns longitudinally reinforced with FRP, however,
strength. In other words, fcc′ and fcu′ had the same value.
reached axial compressive strains of 0.011 and 0.018 in
Therefore, it can be deduced that closed transverse rein-
G-1-120-1.9 and G-3c-80-1.9, respectively, which showed
forcement yield more efficient confinement than C-shaped
that, under good confinement conditions, the FRP bars were
transverse reinforcement because of the material continuity
able to reach high compressive strains.
that eliminates slippage, increasing the lateral confinement
The nominal compressive capacity of the FRP rein-
pressure rather than confinement degradation.
forced columns at peak, considering the gross cross-sec-
tional area Pn, was defined as the sum of the forces carried
Effect of longitudinal reinforcement
by the concrete and the longitudinal reinforcement. Based
The longitudinal reinforcement effect was more
on the elastic theory, the contribution of FRP longitudinal
pronounced in the stress-strain curves based on total
reinforcement bars in compression at peak was calculated
cross-sectional area because the contribution of the longi-
according to the material proprieties given in Table 2. The
tudinal reinforcement was more effective on the pre-peak
authors proposed an equation to calculate the nominal
phase before the activation of the confinement effect.
compressive capacity of the longitudinally and transversally
Figure 12 shows that increasing the FRP longitudinal-
FRP-reinforced columns, which given as follows
reinforcement ratio from 0.8 to 1.0 then 1.9 increased the
load at the peak before activation of confinement. The stress-
Pn = 0.85 × fc′ × (Ag – Afrp) + εco × Efrp × Afrp (SI) (1)
strain curves for these three columns had the same trend. The
three columns failed when the longitudinal bars buckled at
where Ag is gross cross-sectional area of the column; Afrp
nearly the same axial strain. Figure 12 also shows the effect
is cross-sectional area of FRP longitudinal reinforcement;
of longitudinal reinforcement material: Columns G-1-120-
fc′ is concrete compressive strength; εco is concrete strain
1.0 and G-1-120-1.0S had GFRP and steel longitudinal rein-
at peak stress; and Efrp is modulus of elasticity of FRP
forcement, respectively, with the same reinforcement ratio
longitudinal reinforcement.

382 ACI Structural Journal/March-April 2014


Table 4—Prediction of nominal compressive capacity
Column fc′, MPa (ksi) Ac, mm2 (in.2) PExp, kN (lbf) Pn, kN (lbf) Pn/PExp Pc, kN (lbf) PLongi, kN (lbf)
G-1-120-0.8 127,132 (197) 3900 (876,755) 3899 (876,530) 1.00 3782 (850,227) 117 (26,303)
G-1-120-1.0 128,803 (200) 4212 (946,895) 3995 (898,111) 0.95 3832 (861,468) 163 (36,644)
G-1-120-1.9 125,528 (194) 4297 (966,004) 4048 (910,027) 0.94 3734 (839,437) 314 (70,590)
G-1-120-1.0S 124,642 (193) 4272 (960,384) 4260 (957,686) 1.00 3708 (833,592) 552 (124,094)
35 (5.07)
No. 3C-1-67-1.6 127,688 (198) 5159 (1,159,789) 4714 (1,059,749) 0.91 3799 (854,049) 919 (206,599)
No. 3C-1-67-1.0S 125,340 (194) 4660 (1,047,610) 4281 (962,407) 0.92 3729 (838,312) 552 (124,094)
G-3-120-1.9 125,734 (195) 4615 (1,037,493) 4086 (918,569) 0.89 3741 (841,010) 345 (77,559)
C-1-120-1.0S 128,922 (200) 4584 (1,030,524) 4387 (986,237) 0.96 3835 (862,142) 552 (124,094)

Notes: Ac = (Ag – Afrp); Ag is actual total cross-sectional area of considered column.

Fig. 13—Transverse reinforcement layout effect on confined concrete stress-strain response.


Whereas for columns with steel longitudinal reinforce- nally with FRP or steel bars according to Eq. (2) and (1),
ment, the nominal compressive capacity at the peak is given respectively. The results showed that nominal compressive
as follows according to ACI 318 (ACI Committee 318 2008) capacity predictions Pn were conservative and very close to
experimental results, with Pn/PExp ratios varying from 0.89
Pn = 0.85 × fc′ × (Ag – As) + fy × As (2) to 1.00. It is important to note that when Pn differs from
one specimen to another, the load carried by the concrete
where As is the cross-sectional area of longitudinal steel remained similar in all the columns. In other words, the
reinforcement; and fy is the yielding strength of steel difference in Pn is primarily due to the longitudinal rein-
reinforcement. forcement ratio and material, not to concrete strength.
The strength of full-scale plain-concrete columns tested
under concentric compression load is generally lower than Effect of transverse reinforcement
the concrete compressive strength measured on standard The transverse reinforcement restrains the expansion of
150 x 300 mm (6 x 12 in.) cylinders. The 0.85 reduction the concrete core in the column subjected to compressive
factor suggested by ACI 318 (ACI Committee 318 2008) is load and delays its failure. Accordingly, the compressive
mainly attributed to the differences between the reinforced performance of concrete columns depends strongly on the
concrete column and the concrete cylinder regarding concrete transverse reinforcement efficiency. Figure 13 shows the
compressive strength, size, and shape of the element. stress-strain curves of the columns reinforced with different
Table 4 compares the predicted nominal compressive transverse reinforcement layouts to investigate the effect of
capacity at the peak to the experimental results for columns transverse reinforcement configuration, spacing, material,
reinforced transversely with FRP and reinforced longitudi- diameter, and confining volumetric stiffness. The stress-

ACI Structural Journal/March-April 2014 383


strain curves based on the confined concrete core showed GFRP (Table 3) allowed Column G-3-80-1.0S to reach a
that, regardless the transverse reinforcement configura- higher strain than C-3-80-1.0S (0.052 versus 0.034).
tion or material, reducing spacing increases the nominal Regarding transverse reinforcement diameter, C-3-80-1.0S
confined-concrete strength (fcc′/fc′) and changes the behavior outperformed 3C-3-80-1.0S in terms of nominal confined
after this point. With 120 mm (4.7 in.) spacing, a descending concrete strength (1.7 versus 1.4) and ultimate strain (0.034
branch followed the peak stress, while for 80 mm (3.2 in.) versus 0.015). In 3C-3-80-1.0S, however, the longitudinal
spacing, the stress stabilized at a nearly horizontal plateau; bars buckled outside the strain measurement zone.
however, stress increased with the 60 mm (2.4 in.) spacing. Analyzing the results shown in Fig. 13 illustrates the
Configuration 3 proved to be more efficient than Config- effect of the confining volumetric stiffness. Given the same
uration 1, offering higher nominal confined concrete transverse reinforcement layout ρv, the GFRP volumetric
strength (fcc′/fc′) and enhancing the peak stress. In the case of stiffness (ρv × Ef) was far less than that of the CFRP, yet the
120 mm (4.7 in.) spacing, the slope of the descending branch performances were close, which indicates that the config-
following the peak stress was less steep for Configuration uration and spacing are more important parameters than
3 than Configuration 1, resulting in higher ultimate strain modulus of elasticity. Given the same confining volumetric
(almost double) as comparing C-3-120-1.0S and G-3-120- stiffness and transverse reinforcement material, Configu-
1.0S with C-1-120-1.0S and G-1-120-1.0S, respectively ration 3 performed better than Configuration 1 in terms of
(0.039 and 0.041 versus 0.019 and 0.016, respectively). ultimate strain (C-1-60-1.0S and C-3-80-1.0S, as shown in
For columns with 80 mm (3.2 in.) spacing, the stabilization Fig. 13).
plateau was longer in Configuration 3 than Configuration 1.
The ultimate strain increased for the CFRP transversely CONCLUSIONS
reinforced columns from 0.026 to 0.034, corresponding to Failure mechanisms of axially loaded concrete columns
C-1-80-1.0S and C-3-80-1.0S, respectively. The increase of reinforced longitudinally with FRP or steel bars and with
ultimate strain in the GFRP transversely reinforced columns, FRP transverse reinforcement involving different layouts
however, was more than the double comparing G-1-80-1.0S were investigated. Based on the analytical results, the
and G-3-80-1.0S (0.021 and 0.052, respectively). It is clearly following remarks can be made:
shown that the transverse reinforcement spacing and config- 1. The confinement efficiency of closed FRP transverse
uration determined the effectively confined concrete volume, reinforcements cut from continuous square spiral is higher
which increased with closer transverse reinforcement and a than C-shaped type transverse reinforcements.
better distribution of longitudinal bars around the column 2. The ultimate axial strain of columns reinforced longitu-
concrete core. The larger the effectively confined concrete dinally with FRP is almost 30% lower than those reinforced
volume, the higher the confinement efficiency. In addition, with the same volume of steel.
transverse reinforcement spacing controlled the buckling of 3. The ultimate axial compressive strain for columns rein-
the longitudinal bars by reducing their slenderness ratio. forced longitudinally and transversally with FRP can reach
The effect of transverse reinforcement material was a value on the same order of magnitude as the FRP ultimate
related to columns mode of failure, which is dependent on tensile strain of the longitudinal bars under good confine-
the configuration. Figure 13(b), (c), and (d) showed that, in ment conditions.
the case of columns with Configuration 1 (C-1-120-1.0S, 4. The contribution of FRP longitudinal reinforcement
G-1-120-1.0S, C-1-80-1.0S, and G-1-80-1.0S) that failed in concrete columns subjected to axial concentric loading
because of excessive longitudinal bars buckling, CFRP should not be neglected. A proposed equation based on
transverse reinforcement enabled the columns to attain elastic theory yields good predictions compared with labo-
higher nominal confined concrete strength (fcc′/fc′) than those ratory test data.
reinforced with GFRP transverse reinforcement. The stress- 5. FRP transverse reinforcement configuration and
strain curves of both materials, however, followed the same spacing are the most important parameters (compared with
trend. The modulus of elasticity was determinant: given confinement provided by concrete cover) affecting confining
the same layout, the CFRP transverse reinforcements were efficiency in internally reinforced concrete columns under
stiffer than the GFRP ones. The stiffer transverse reinforce- axial loading.
ment tended to open less, thereby better limiting the buck- 6. In the case of large spacing with low volumetric ratio,
ling of longitudinal bars. Two different spacing related cases CFRP transverse reinforcement performed significantly
were observed in columns with Configuration 3. The first better than GFRP. Increasing the volumetric ratio while
relates to 120 mm (4.7 in.) spacing, which was wide enough reducing spacing will eliminate the effect of material stiff-
to enable significant bar buckling and the development of ness. In such cases, the GFRP transverse reinforcements are
a localized plastic hinge, as illustrated in Fig. 9(a) and (b). more cost effective.
The columns reinforced transversely with CFRP and GFRP 7. Columns internally reinforced with a combination of
behaved identically. In the second case, the stirrups spacing steel longitudinal bars and FRP transverse reinforcements
of 80 mm (3.2 in.) prevented excessive bar buckling and exhibit good gains in terms of compressive strength and
the development of a localized plastic hinge. In this case, ultimate axial strain. Nonetheless, the use of FRP transverse
the CFRP stirrups allowed the column to achieve higher reinforcement should still improve corrosion resistance of
nominal confined concrete strength than GFRP stirrups (1.7 a column by adding an extra 10 mm (0.4 in.) of cover to
versus 1.6). Conversely, the larger ultimate elongation of the steel.

384 ACI Structural Journal/March-April 2014


8. The presented study showed the applicability of exclu- CSA S807, 2010, “Specification for Fibre-Reinforced Polymers,” Cana-
dian Standards Association, Mississauga, ON, Canada, 44 pp.
sively reinforcing the columns with FRP and subjected to Cusson, D., and Paultre, P., 1994, “High Strength Concrete Columns
concentric load. Further research elaboration is necessary Confined by Rectangular Ties,” Journal of Structural Engineering, ASCE,
to investigate the behavior of FRP reinforced columns V. 120, No. 3, Mar., pp. 783-804.
Federation Internationale de Béton (FIB), 2007, “FRP Reinforcement in
loaded laterally or subjected to load combination (axially RC Structures,” Task Group 9.3, Lausanne, Switzerland, 157 pp.
and laterally). Harries, K. A., and Kharel, G., 2003, “Experimental Investigation of the
Behavior of Variably Confined Concrete,” Cement and Concrete Research,
V. 33, pp. 873-880.
AUTHOR BIOS ISIS Canada Research Network, 2007, “Reinforcing Concrete Structures
Hany Tobbi is a Doctoral Candidate in the Department of Civil Engi-
with Fibre Reinforced Polymers,” ISIS Design Manual No. 3, ISIS Canada
neering at the University of Sherbrooke, Sherbrooke, QC, Canada. He
Research Network, 151 pp.
received his BSc from the University of Mentouri, Constantine, Algeria, and
Mander, J. B.; Preistley, M. J. N.; and Park, R., 1988a, “Theoretical
his MSc from the University of Claude Bernard, Lyon, France. His research
Stress-Strain Model for Confined Concrete,” Journal of Structural Engi-
interests include structural analysis, design, and testing of concrete struc-
neering, ASCE, V. 114, No. 8, pp. 1804-1826.
tures reinforced with fiber-reinforced polymers.
Mander, J. B.; Preistley, M. J. N.; and Park, R., 1988b, “Observed
Stress-Strain Behaviour of Confined Concrete,” Journal of Structural Engi-
Ahmed Sabry Farghaly is a Postdoctoral Fellow in the Department of
neering, ASCE, V. 114, No. 8, pp. 1827-1849.
Civil Engineering at the University of Sherbrooke, and Associate Professor
Nanni, A., 1993, “Flexural Behavior and Design of RC Members Using
in the Department of Civil Engineering, Assiut University, Assiut, Egypt.
FRP Reinforcement,” Journal of Structural Engineering, ASCE, V. 119,
His research interests include nonlinear analysis of reinforced concrete
No. 11, pp. 3344-3359.
structures, and behavior of structural concrete reinforced with fiber-rein-
Pultrall, Inc., 2009, V-ROD Composite Reinforcing Rods Technical Data
forced polymers.
Sheet, Thetford Mines, Canada, www.pultrall.com.
Richart, F. E.; Brandtzaeg, A.; and Brown, R. L., 1928, “A Study of
Brahim Benmokrane, FACI, is an NSERC Research Chair in FRP Rein-
the Failure of Concrete under Combined Compressive Stresses,” Bulletin
forcement for Concrete Infrastructures and Tier-1 Canada Research Chair
No. 185, Engineering Experimental Station, University of Illinois, Urbana,
Professor in Advanced Composite Materials for Civil Structures in the
IL, 104 pp.
Department of Civil Engineering at the University of Sherbrooke. He is a
Richart, F. E.; Brandtzaeg, A.; and Brown, R. L., 1929, “The Failure of
member of ACI Committee 440, Fiber-Reinforced Polymer Reinforcement.
Plain and Spirally Reinforced Concrete in Compression,” Bulletin No. 190,
Engineering Experimental Station, University of Illinois, Urbana, IL, 74 pp.
ACKNOWLEDGMENTS Saatcioglu, M., and Razvi, S. R., 1992, “Strength and Ductility of
The authors would like to express their special thanks and gratitude to the Confined Concrete,” Journal of Structural Engineering, ASCE, V. 118,
Natural Science and Engineering Research Council of Canada (NSERC), No. 6, pp. 1590-1607.
the Fonds québécois de la recherche sur la nature et les technologies Saatcioglu, M.; Salamat, A. H.; and Razvi, S. R., 1995, “Confined
(FQRNT), the Canadian Foundation for Innovation (FCI), and the technical Columns under Eccentric Loading,” Journal of Structural Engineering,
staff of the structural lab in the Department of Civil Engineering at the ASCE, V. 121, No. 11, pp. 1547-1556.
University of Sherbrooke. Sheikh, S. A., 1982, “A Comparative Study of Confinement Models,”
ACI Journal, V. 79, No. 4, July-Aug., pp. 296-306.
Sheikh, S. A., and Uzumeri, S. M., 1980, “Strength and Ductility of Tied
REFERENCES Concrete Columns,” Journal of the Structural Division, ASCE, V. 106,
ACI Committee 318, 2008, “Building Code Requirements for Structural
No. 5, pp. 1079-1112.
Concrete (ACI 318-08) and Commentary,” American Concrete Institute,
Sheikh, S. A., and Uzumeri, S. M., 1982, “Analytical Model for Concrete
Farmington Hills, MI, 473 pp.
Confinement in Tied Columns,” Journal of the Structural Division, ASCE,
ACI Committee 440, 2004, “Guide Test Methods for Fiber-Reinforced
V. 108, No. 12, pp. 2703-2722.
Polymers (FRPs) for Reinforcing or Strengthening Concrete Structures
Teng, J. G.; Chen, J. F.; Smith, S. T.; and Lam, L., 2002, FRP Strength-
(ACI 440.3R-04),” American Concrete Institute, Farmington Hills, MI, 40 pp.
ened RC Structures, John Wiley & Sons, Ltd., Hoboken, NJ, 266 pp.
ACI Committee 440, 2006, “Guide for the Design and Construction of
Tobbi, H.; Farghaly, A. S.; and Benmokrane, B., 2012, “Concrete
Structural Concrete Reinforced with FRP Bars (ACI 440.1R-06),” Amer-
Columns Reinforced Longitudinally and Transversally with GFRP Bars,”
ican Concrete Institute, Farmington Hills, MI, 44 pp.
ACI Structural Journal, V. 109, No. 4, July-Aug., pp. 551-558.
ACI Committee 440, 2007, “Report on Fiber-Reinforced Polymer (FRP)
Watson, S.; Zahn, F. A.; and Park, R., 1994, “Confining Reinforcement
Reinforcement Concrete Structures (ACI 440R-07),” American Concrete
for Concrete Columns,” Journal of Structural Engineering, ASCE, V. 120,
Institute, Farmington Hills, MI, 100 pp.
No. 6, pp. 1798-1824.
ACI Committee 440, 2008, “Specification for Carbon and Glass
Fiber-Reinforced Polymer Bar Materials for Concrete Reinforcement (ACI
440.6M-08),” American Concrete Institute, Farmington Hills, MI, 6 pp.
CSA S806, 2012, “Design and Construction of Building Components
with Fiber-Reinforced Polymers,” Canadian Standards Association, Missis-
sauga, ON, Canada, 177 pp.

ACI Structural Journal/March-April 2014 385


NOTES:

386 ACI Structural Journal/March-April 2014


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 111-S34

Repair of Prestressed Concrete Beams with Damaged


Steel Tendons Using Post-Tensioned Carbon Fiber-
Reinforced Polymer Rods
by Clayton A. Burningham, Chris P. Pantelides, and Lawrence D. Reaveley

Research implementing unibody clamp anchors and a simple Despite its historic use, however, exposed steel is susceptible
mechanical stressing device to post-tension external, unbonded to corrosion, limiting its useful lifespan and requiring exten-
carbon fiber-reinforced polymer (CFRP) rods is presented. The sive protection from deicing salt and moisture.
experiments described in the paper concern three prestressed The limitations of steel tendons can be overcome in
concrete beams: one was used as the control beam and the other
external post-tensioning applications by the use of fiber-rein-
two were damaged. Damage consisted of cracked concrete that
forced polymer (FRP) materials. FRP materials are advanta-
was removed and internal steel tendons that were cut to simulate
vehicle collision, corrosion, or both. The repair system was then geous because of their corrosion resistance and high specific
applied to the two damaged concrete beams. The CFRP repair strength. Additionally, the use of FRP materials is becoming
system performed well, increasing the ultimate strength and flex- increasingly attractive as the price of FRP composites
ural capacity of the damaged beams to meet or exceed the strength decreases. Several studies have shown that post-tensioned
capacity of the control. An analytical model considering the tendon FRP tendons can contribute to flexural strength in new
stress at ultimate and the distribution of internal forces was devel- construction or for strengthening (Abdel Aziz et al. 2005;
oped to explore design recommendations for the use of the unibody El-Hacha and Elbadry 2006; Täljsten and Nordin 2007);
clamp anchors and stressing device for post-tensioning CFRP rods. however, few studies have shown the usefulness of post-ten-
sioned FRP tendons in flexural repair and retrofit applica-
Keywords: beams; carbon fiber-reinforced polymer (CFRP); post-ten-
sioning; prestressed concrete; repair; retrofit. tions (Elrefai et al. 2007). As a result, additional research is
required to investigate the suitability of post-tensioned FRP
INTRODUCTION tendons for the repair of severe flexural damage.
Many bridges in the Unites States are approaching the Widespread use of FRP tendons in post-tensioning appli-
end of their design life, and some bridges are showing signs cations has been slow because of the difficulty in devel-
of aging and damage such as corrosion of steel reinforce- oping an effective tendon anchor. Research has produced
ment, large cracks, and missing concrete cover. Damage to a unibody clamp anchor and mechanical stressing device
concrete cover and internal steel prestressing tendons can for use in post-tensioning carbon FRP (CFRP) rods (Burn-
occur when large vehicles attempt to pass under a bridge ingham 2011). The clamp anchors are machined from a
without adequate clearance. Vehicular impact can fracture single piece of steel, and the clamping force is provided by
the concrete cover, expose the internal steel prestressing high-strength bolts. Further research is needed to analyze
tendons, and/or sever all or part of the outer steel prestressing the effectiveness of the complete post-tensioning system
tendons. Even if the tendons are not severed, removal of the consisting of the CFRP rods, unibody clamp anchors, and
protective concrete cover accelerates the corrosion process. mechanical stressing device when applied to prestressed
Additionally, cracking from overloading or fatigue could concrete members.
facilitate corrosion of internal steel prestressing tendons. The research in this paper is concerned with applying
Damage to internal steel prestressing tendons decreases flex- CFRP rods, unibody clamp anchors, and the aforementioned
ural capacity, and bridges exhibiting these symptoms could mechanical stressing device as a complete FRP strength-
be in critical need of replacement, repair, or strengthening. ening system for the repair of damaged prestressed concrete
Typically, girder replacement is expensive, time beams. In the present research, the unibody clamp anchors
consuming, and disruptive; therefore, repair or retrofit is often were fabricated using mild steel. In actual implementation,
the preferred option. One system used for repair applications the anchor and stressing device might need to be manufac-
is external post-tensioning. This repair method not only tured using stainless steel or other corrosion-resistant steel.
restores flexural capacity, but can also mitigate the demands The specific damage considered during this research was
of an increase in service load and help with serviceability damage resulting from impact with vehicles passing under-
considerations such as deflection. Thus, external post-ten- neath a bridge without adequate clearance. Such impact
sioning is an excellent option for repairing concrete bridge could result in severed internal steel prestressing tendons
girders with damage to internal steel prestressing tendons. ACI Structural Journal, V. 111, No. 2, March-April 2014.
Traditionally, external post-tensioning has been imple- MS No. S-2012-135.R2, doi:10.14359.51686529, was received October 18, 2012,
and reviewed under Institute publication policies. Copyright © 2014, American
mented with high-strength steel tendons because of low Concrete Institute. All rights reserved, including the making of copies unless
permission is obtained from the copyright proprietors. Pertinent discussion including
material cost, material availability, and ease of installation. author’s closure, if any, will be published ten months from this journal’s date if the
discussion is received within four months of the paper’s print publication.

ACI Structural Journal/March-April 2014 387


or removal of concrete cover and subsequent corrosion of ating the ultimate stress of unbonded post-tensioned CFRP
internal steel prestressing tendons. An illustrative example rods. The CFRP repair system implemented in this research
of impact damage observed on an actual prestressed concrete could facilitate the acceptance of CFRP post-tensioning
bridge girder can be seen in Fig. 1. systems by the construction industry.
The paper includes details of testing methods, specimen
design and fabrication, experimental design, and analysis of EXPERIMENTAL INVESTIGATION
results. Additionally, the methods used for collecting data
during laboratory testing as well as details pertaining to the Specimen fabrication
performance and effectiveness of the FRP repair system Three prestressed concrete (PC) beam specimens were
and its application are provided, with specific focus on the designed and fabricated for testing. The three beams (Spec-
performance of the CFRP tendons and their ability to aid in imens P2, RP1, and RP3, with “R” indicating a specimen to
the repair of damaged beams. An analytical model consid- which the repair system was applied) were manufactured by
ering the tendon stress at ultimate and conventional beam a local PCI-certified precast/prestressed concrete company.
theory is presented to explore design recommendations for The precast beams measured 12 in. (305 mm) wide x 20 in.
the use of the unibody clamp anchors and stressing device. (508 mm) tall x 15 ft (4.57 m) long, and each prestressed
beam had three 1/2 in. (13 mm) seven-wire low-relax-
RESEARCH SIGNIFICANCE ation prestressing steel strands with an ultimate strength of
Previous research on using external post-tensioned CFRP 270 ksi (1862 MPa). The beams were also reinforced with
tendons for repair of damaged concrete beams has been two No. 5 (16 mm) mild steel bars in the tension zone, and
limited. This research presents the implementation of newly two No. 5 (16 mm) mild steel bars in the compression zone.
developed unibody clamp anchors and a simple mechan- The prestressed beams had No. 3 (10 mm) stirrups placed at
ical stressing device for the repair of damaged prestressed 12 in. (305 mm) on center. The beam dimensions and loca-
concrete beams with post-tensioned CFRP rods. In addition, tion of internal reinforcement are shown in Fig. 2.
the paper validates equations from the literature for evalu-
Experimental design
All three beams were tested and subjected to initial damage
using a four point loading system to induce tensile cracking.
The same setup was used to test the beams to failure. A
hydraulic actuator with a 500 kip (2220 kN) inline load cell
and a steel spreader beam were used to apply a two-point
load, spaced 30 in. (762 mm) apart, to the top of the speci-
mens, as shown in Fig. 3. The specimens were tested with
an unbraced length of 13 ft, 8 in. (4.17 m) and had a depth
of 20 in. (508 mm), giving a shear span-depth ratio (a/d)
of 3.35.

Material properties
The materials used in this research are typical of construc-
tion in the United States. All steel reinforcing bars used in the
Fig. 1—Girder damage from vehicle impact. fabrication of the specimens had a nominal tensile strength

Fig. 2—Reinforcement layout for Specimens P2, RP1, and RP3. (Note: 1 in. = 25.4 mm; 1 ft = 0.305 m.)

388 ACI Structural Journal/March-April 2014


Fig. 4—Damaged outer steel tendon in Specimen RP1 (three
wires cut) and RP3 (seven wires cut).
In addition, the rate of displacement was held constant at
Fig. 3—Test setup. (Note: 1 in. = 25.4 mm; 1 ft = 0.305 m.) 0.0625 in./min (1.59 mm/min) throughout the test. All spec-
imens were subjected to the same loading protocol, with
of 60 ksi (414 MPa). The CFRP rods used in this research termination of loading dependent upon the level of cracking.
had the following properties as provided by the manufac- The vertical deflection at midspan was limited to 0.375 in.
turer: rod diameter = 3/8 in. (9.53 mm), tensile strength (9.5 mm), and the cracks were of a hairline width mainly in
= 250 ksi (1724 MPa), tensile modulus = 22,500 ksi (155 the constant moment region at a spacing of 12 in. (305 mm).
GPa), and elongation at break = 1.11%. In separate tensile Additional damage was inflicted on Specimens RP1 and
tests of CFRP rods from the same lot as the ones used in RP3 to simulate damage to internal steel prestressing tendons
this research, carried out using unibody clamp anchors, the from vehicle collision, subsequent corrosion, or both. An
average measured tensile strength was 308 ksi (2124 MPa), area approximately 8 in. (203 mm) long with a depth equal to
and the average ultimate strain measured was 1.37%. The the concrete cover of the prestressing tendons was removed
internal steel prestressing tendons were low relaxation from both Specimen RP1 and RP3 to expose an outer seven-
1/2 in. (13 mm) diameter seven-wire strands with a nominal wire steel prestressing strand within the constant moment
ultimate strength of 270 ksi (1862 MPa). Concrete cylinder region. For Specimen RP1, three of the seven wires in this
tests performed at 7 days after casting of the steam cured strand were cut—leaving two intact seven-wire strands and
prestressed concrete beams gave an average compressive one four-wire damaged strand. For Specimen RP3, all seven
concrete strength of 7.0 ksi (48 MPa), and at the time of wires of an outer steel prestressing strand were cut to simulate
specimen testing, the concrete had an average compressive severe damage, leaving two intact seven-wire strands. These
strength of 10.0 ksi (69 MPa) based on compression tests of cuts, seen in Fig. 4, simulated partial or complete severing of
4 in. (102 mm) diameter by 8 in. (204 mm) high cylinders. the exterior tendon on impact in an exterior girder or corro-
sion of an exterior tendon due to loss of concrete cover and
Testing methods subsequent exposure to the elements.
Load testing was carried out in three phases: damage, FRP repair—After simulating damage to internal steel
repair, and failure. First, loading was used to introduce tensile prestressing tendons in Specimens RP1 and RP3, the beams
cracks, which could lead to accelerated corrosion of internal were repaired with external post-tensioned CFRP rods. The
steel prestressing tendons. Additionally, Specimens RP1 and unibody clamp anchors and mechanical stressing device
RP3 were damaged with respect to the internal prestressing seen in Fig. 5 were used to introduce the post-tensioning
steel—to simulate impact damage. Subsequently, Spec- force. Specimens RP1 and RP3 were repaired with two rods,
imens RP1 and RP3 were repaired with external post-ten- one on each side of the beam along the beam length at a
sioned CFRP rods. Finally, all three specimens were loaded depth of 15 in. (381 mm) from the top compression fiber.
monotonically to failure. The CFRP rods were post-tensioned to a strain of approx-
Damage loading—Damage loading applied to the spec- imately 0.485% measured using strain gauges attached to
imens consisted of downward half-cycles to induce tensile the rod at midspan, producing a calculated design post-ten-
cracking. The loading was displacement controlled to sioning force of 12 kip (53.4 kN) in each rod. ACI 440.4R
avoid catastrophic failure and subsequent loss of the spec- (ACI Committee 440 2004) recommends initial jacking
imens. Displacement half-cycles were applied in incre- stresses of 40 to 65% of the design ultimate tensile strength
ments of 0.0625 in. (1.59 mm), with the amplitude of each of the prestressed FRP tendon and anchorage system. In
successive half-cycle increasing by 0.0625 in. (1.59 mm). the present case, the amount of initial jacking of the CFRP

ACI Structural Journal/March-April 2014 389


Fig. 6—Control Specimen P2 at failure; grid = 4 in.
Fig. 5—End view of stressing system showing clamp anchors (102 mm).
and stressing device.
rods was approximately 44% of the design ultimate tensile
strength based on the manufacturer’s specification.
The novel mechanical stressing device implemented
in this research consists of a slotted square HSS section
running perpendicular to the beam length at both ends of
the beam. The slots in the HSS section allow the tendons to
pass through the slots such that the unibody clamp anchors
make contact with the back side of the tube. On the stressing
end of the beam, two sleeve nuts are positioned on top, and
two on the bottom of the HSS section. The sleeve nuts run
parallel with the beam, and tendon stressing occurs when
1.0 in. (25 mm) diameter bolts are screwed into the sleeve
nuts; the bolts react against the beam end, moving the HSS
section back to stress the tendons. Tightening the stressing
bolts in an alternating star pattern ensures the tendons are
stressed with controlled increments of tightening. More
details of the unibody clamp anchor and stressing device are
provided by Burningham (2011).
Loading to failure—After repairing Specimens RP1 and
RP3 with external post-tensioned CFRP rods, the specimens
were loaded to failure, with Specimen P2 as the control
specimen. The displacement controlled loading to failure
was monotonic at a rate of 0.0625 in./min (1.59 mm/min).
Fig. 7—Repaired Specimen RP1 at imminent failure:
The test was stopped at imminent failure of the specimen,
(a) front; and (b) back; grid = 4 in. (102 mm).
as measured by a 20% decrease from the maximum load or
failure of the external CFRP rods, whichever occurred first. measured in units of microstrain, and all LVDT readings
were measured within 0.001 in. (0.025 mm).
EXPERIMENTAL RESULTS
Specimen data analysis
Instrumentation and data collection methods No anchor slippage was observed in any of the CFRP rods
Instrumentation consisted of strain gauges and linear vari- during testing of the specimens to failure. The lack of slip
able differential transformers (LVDTs). The specimens were demonstrates that the anchors work as designed. This was
instrumented with three LVDTs, one 54.5 in. (1.38 m) from confirmed in the laboratory in CFRP rod tensile tests using
either end and one at midspan as shown in Fig. 3; these were the unibody anchors according to ACI 440.3R guidelines
attached to the bottom of the beam to measure the deflected (ACI Committee 440 2012).
shape under load. Concrete strain gauges placed at 69.5 in. During testing, Specimen P2 (control) failed due to
(1.77 m) from each end and at midspan on the top face of concrete compressive failure at the center, after flexural
the beams measured the concrete compressive strain. Strain cracks had developed; crack spacing in the constant moment
gauges were also applied to each CFRP rod at midspan as region was 6 in. (152 mm) and the maximum crack width
shown in Fig. 3; they were used to measure strain at initial was 0.08 in. (2 mm), as shown in Fig. 6. Specimen RP1
prestress and throughout the tests until failure of Speci- (repaired) failed due to concrete compression failure accom-
mens RP1 and RP3. All measurements were collected by panied by flexural cracks, after rupture of the external CFRP
an electronic data acquisition system at a sampling rate of rods; crack spacing in the constant moment region was
two data points per second. All strain gauge readings were 6 in. (152 mm), and the maximum crack width was 0.06 in.

390 ACI Structural Journal/March-April 2014


Fig. 9—Concrete compressive strain at midspan versus
applied load to failure.

Fig. 8—Repaired Specimen RP3 at failure: (a) front;


(b) back; grid = 4 in. (102 mm).
(1.5 mm), as shown in Fig. 7. Specimen RP3 (repaired)
failed due to concrete compression failure at the center;
flexural crack spacing in the constant moment region was Fig. 10—CFRP rod strain at midspan versus applied load
4 in. (102 mm), and the maximum crack width was 0.07 in. to failure.
(1.8 mm). In addition, Specimen RP3 developed several
flexural tension cracks around the area of the cut tendon, and tively. Because the measured ultimate strain of the rods in
the top compression mild steel bars buckled subsequent to axial tension is 1.37, the most likely explanation for rupture
concrete crushing, as shown in Fig. 8. of the rods at lower strains is eccentric bending due to lack
For all three specimens, the highest compressive concrete of rotation of the anchor at the supports; this caused stress
strain was observed near midspan, as seen in Fig. 9, which concentrations at a point other than where the strain gauges
shows the change in concrete compressive strain at midspan were located at midspan. One location of stress concentra-
as a function of applied load. From Fig. 9 it can be seen tions is near the clamp anchors at the beam ends due to rota-
that Specimens P2 and RP3 experienced maximum concrete tion of the latter at large deflections. Figure 11 shows the two
strains just greater than 0.30%. Specimen RP1 also started modes of failure observed: straw broom and splitting failure;
failing in compression, but the test was terminated prema- the former failure was observed near the clamp anchors, and
turely at a smaller deflection than the other two tests with the the latter near midspan. In actual applications, the use of a
maximum concrete strain reaching 0.27%. These numerical rotating rocker at both anchorage devices is desirable to keep
data correlate well with the failure mode visually observed the rod from experiencing flexural bending stresses.
in Specimens P2 and RP3 and seen in Fig. 6 and 8, respec- The CFRP repair system was successful—both Specimens
tively. Figure 7 shows that Specimen RP1 experienced severe RP1 and RP3 exhibited an increase in ultimate capacity and
cracking damage with significant flexural cracks reaching benefited from application of the CFRP repair system. The
the top of the beam, and imminent concrete compression ultimate load for Specimens P2 (control), RP1 (repaired),
failure, similar to Specimen P2. and RP3 (repaired) was 104, 112, and 102 kip (463, 498,
The change in CFRP rod strain at midspan as a function of and 454 kN), respectively. For Specimen RP1, the ultimate
applied load for Specimens RP1 and RP3 is shown in Fig. 10. load corresponds to an increase of approximately 7.7% in
An increase in strain of the CFRP rods for Specimens RP1 ultimate capacity from the use of external post-tensioned
and RP3 was observed during testing of the specimens to CFRP rods. It should be remembered, however, that Spec-
failure. The initial strain in the rods from the post-tensioning imen RP1 (repaired) had two intact seven-wire strands and
application was 0.485%. At failure of the external CFRP one four-wire strand (three wires were cut), whereas Spec-
tendons, the average maximum strain in the CFRP rods was imen P2 (control) had three intact seven-wire strands. There-
0.750 and 0.814% for Specimens RP1 and RP3, respec- fore, application of the theoretical capacity of Specimen
RP1 (repaired) based on the cut wires produces an effec-

ACI Structural Journal/March-April 2014 391


Fig. 12—Applied load versus midspan deflection to failure.

Fig. 11—Failure of CFRP rods: (a) straw broom; and


(b) splitting.
tive increase in ultimate capacity of 20.6% from the use of
external post-tensioned CFRP rods. In addition, although a
third of the prestressing force was removed from Specimen
RP3, the repair with the external post-tensioned CFRP rods
produced load-deflection behavior virtually identical to that
of the control specimen. This identical behavior implies
restoration of the girder and an effective increase in ultimate
capacity of 31.1% from the use of external post-tensioned
CFRP rods.
Deformability and ductility—The similarities in the
Fig. 13—Member ductility of concrete beams prestressed by
performance of Specimens P2 (control) and RP3 (repaired)
FRP by Abdelrahman et al. (1995).
are shown in Fig. 12. From Fig. 12, it can be seen that failure
of the CFRP rods for Specimen RP1 occurred at a deflection ductility for FRP reinforced beams is the method developed
about eight times greater than L/800 (0.21 in. or 5.21 mm), by Abdelrahman et al. (1995). As shown in Fig. 13, ductility
the maximum allowable design deflection under service live is defined in terms of the equivalent displacement of the
loads for concrete bridge construction (AASHTO 2009). uncracked section and the displacement at ultimate and is
Therefore, although failure of the post-tensioned CFRP rods given by the following expression
was brittle, failure of the beams occurred at a deflection much
greater than any expected service load deflections. After ∆2
failure of the CFRP rods, Specimen RP1 (repaired) exhib- µ∆ = (1)
∆1
ited a residual strength of approximately 97 kip (431 kN).
This residual strength is evidence that complete catastrophic
failure of the beam did not occur because of the reserve where Δ1 is the displacement of the uncracked section at a
capacity and ductility of the original system. Additionally, load equal to the ultimate load; and Δ2 is the displacement at
it can be concluded from Fig. 12 that the residual strength failure. Using Eq. (1) and Fig. 12, the ductility of the control
in Specimen RP1 (repaired) after failure of the CFRP rods beam Specimen P2 is obtained as 8.1, and that of the repaired
suggests that at large deflections, it would fail similarly to beams as 7.7 for RP3 and 7.3 for RP1. It is observed that the
Specimen P2 (control)—from concrete compressive failure; repaired beams are practically as ductile as the control beam.
the test for Specimen RP1 was terminated prematurely at a It should be noted that Specimen RP1 would have achieved
smaller deflection. greater ductility if the test had not been terminated.
Another measure of performance is the ductility of the
repaired beams compared with the control beam. Ductility ANALYTICAL INVESTIGATION
is provided by the mild steel present in the tension and Conventional beam theory can be used to predict the
compression zone of all three beams, as shown in Fig. 2. ultimate load of the specimens tested in this research. The
CFRP rods are brittle, and even though the amount of mild nominal theoretical capacity of control Specimen P2 was
steel was the same, the amount of prestressing steel was calculated to be 73 kip (330 kN). Compared with the actual
less for the repaired beams; therefore, it is interesting to ultimate load of 104 kip (463 kN), the ratio of actual to theo-
compare the ductility of the three beams. One definition of retical ultimate load is 1.42, indicating that the theoretical
prediction is in good agreement with the actual value, and the

392 ACI Structural Journal/March-April 2014


Table 1—Data for calculation of theoretical values at ultimate
Specimen fp_CFRP, ksi (MPa) cu, in. (mm) Aps, in.2 (mm2) a, in. (mm) Mu, kip-ft (kN-m)
RP1 (repaired) 155 (1069) 2.51 (63.8) 0.393 (254) 1.63 (41.4) 220 (298)
PR3 (repaired) 162 (1117) 2.22 (56.4) 0.306 (197) 1.44 (36.6) 196 (266)

design is conservative. To determine the theoretical capacity steel prestressing strands; As is the area of tensile mild steel
of Specimens RP1 and RP3, the stress in the CFRP rods at reinforcement, 0.62 in.2 (400 mm2); b is the beam width,
ultimate should first be determined. Previous research for 12 in. (305 mm); fc′ is the compressive strength of concrete,
unbonded steel tendons has shown that strain compatibility 10.0 ksi (69 MPa); fps is the prestressing force in the internal
can be used to analyze the tendons as if they were bonded, steel strands, 243 ksi (1.68 GPa); fy is the yield stress of
and then apply a strain reduction factor to account for the mild steel reinforcement, 60 ksi (414 MPa); and β1 is 0.65.
tendons being unbonded (Naaman and Alkhairi 1991). It has The use of Eq. (4) produces cu values of 2.51 and 2.22 in.
been suggested that this method of using a strain reduction (63.8 and 56.4 mm) for Specimens RP1 and RP3, respec-
factor could also be applied to FRP tendons (Naaman et al. tively. The resulting CFRP rod ultimate stresses predicted
2002; ACI Committee 440 2004); the stress at ultimate in by Eq. (2) are 155 ksi (1069 MPa) for Specimen RP1, and
unbonded FRP tendons is given as 162 ksi (1120 MPa) for Specimen RP3, which are conserva-
tive compared with the actual values of measured ultimate
d  CFRP rod stress (from strain gauges on the rods) of 169 and
f p _ CFRP = f pe _ CFRP + Ωu ECFRP e cu  CFRP − 1 (2) 183 ksi (1163 and 1263 MPa), respectively. The analytical
 cu 
results for ultimate stress are summarized in Table 1.
An alternative strain reduction factor for external,
where cu is the depth to the neutral axis at ultimate; dCFRP is unbonded steel tendons has also been developed, and it had
the depth to the CFRP rods, 15 in. (381 mm); ECFRP is the been recommended for FRP tendons (Aravinthan et al. 1997;
modulus of elasticity of the CFRP rods, 22,500 ksi (155 GPa); ACI Committee 440 2004). This alternative strain reduction
fp_CFRP is the stress in the CFRP rods at ultimate; fpe_CFRP is the factor is given as
effective prestress in the CFRP rods, 109 ksi (752 MPa); Ωu
is the strain reduction factor; and εcu is the failure strain of 0.21 A 
concrete in compression, 0.003 in./in. (mm/mm). Ωu = + 0.04  CFRP int  + 0.04 (5)
L / dCFRP  ACFRP tot 
Suggested values for the strain reduction factor depend on
the type of loading. The research presented in this paper is
best described as center point loading because the distance where ACFRP int is the area of internal CFRP tendons; and
between loading points was only 30 in. (762 mm) compared ACFRP tot is the total area of CFRP tendons. Additionally, to
with an unbraced length of 13ft, 8 in. (4.17 m). The strain account for the change in an external tendon’s eccentricity,
reduction factor—standardized to likely produce a conser- the effective CFRP tendon depth can be found using
vative predicted value—for center point loading is given as
de _ CFRP = dCFRP Rd (6)

1.5
Ωu = (3)
L / dCFRP where Rd is the depth reduction factor given as

where L is the unbraced length of the beam, 13 ft, 8 in.  L  S 


(4.17 m) (Naaman et al. 2002; ACI Committee 440 2004). Rd = 1.14 − 0.005   − 0.19  d  ≤ 1.0 (7)
 dCFRP   L
The use of Eq. (3) for specimens in this research results in
a strain reduction factor of 0.137. Additionally, appropriate
values of cu can be calculated from the following equation with the deviator spacing Sd equal to the span L for speci-
mens in this research because no deviators were used. For
the case of specimens in this research, the value of Rd is
− B + B2 − 4 AC
cu = (4a) 0.895. The predicted ultimate CFRP tendon stresses from
2A Eq. (2) using the alternative strain reduction factor from
Eq. (5) are 135 and 138 ksi (931 and 951 MPa) for Speci-
mens RP1 and RP3, respectively. Similar to the results from
A = 0.85 fc′bb1 the application of the strain reduction factor of Eq. (3), these
( )
B = ACFRP ECFRP e cu Ωu − f pe _ CFRP − As f y − Aps f ps (4b)
predictions are conservative. A comparison of the applica-
bility of the two strain reduction factors can be made when
C = − ACFRP ECFRP e cu Ωu dCFRP considering the actual and predicted ultimate CFRP tendon

stresses for Specimens RP1 and RP3. The measured CFRP
In the aforementioned expressions, ACFRP is the area of tendon stresses, the predicted, and the error are shown in
CFRP rods, 0.22 in.2 (142 mm2); Aps is the area of internal Table 2. Although both strain reduction factors produced

ACI Structural Journal/March-April 2014 393


Table 2—Measured and predicted ultimate tendon stresses
Predicted CFRP tendon stress
Measured CFRP tendon Predicted CFRP tendon stress at Error in prediction at ultimate from Eq. (5), ksi Error in prediction
Specimen stress at ultimate, ksi (MPa) ultimate from Eq. (3), ksi (MPa) from Eq. (3), % (MPa) from Eq. (5), %
RP1 (repaired) 169 (1163) 155 (1069) 8 135 (931) 18

PR3 (repaired) 183 (1263) 162 (1120) 11 138 (951) 24

conservative ultimate stress predictions, the strain reduction Table 3—Experimental and theoretical
factor from Eq. (3) results in the smallest error, indicating ultimate loads
that it works best for specimens in this research.
Theoretical Ratio of
Conventional beam theory leads to the following equation ultimate load experimental
for ultimate moment capacity Experimental ulti- from Eq. (3), to theoretical
Specimen mate load, kip (kN) kip (kN) ultimate load
 a  a P2 (control) 104 (463) 73 (325) 1.42
Mu = Aps f ps  d ps −  + As f y  d − 
 2  2 RP1 (repaired) 112 (498) 79 (351) 1.42
 a (8)
+ ACFRP f p _ CFRP  dCFRP −  PR3 (repaired) 102 (454) 70 (311) 1.46
 2

and RP3 showed an effective increase in ultimate strength of
where Mu is the ultimate moment capacity; a is the depth of 20.6 and 31.1%, respectively, with respect to the damaged
the equivalent compression stress block equal to β1cu; and d condition. This increase in ultimate strength of Specimens
is the depth to the mild steel reinforcement, 15 in. (381 mm). RP1 (repaired) and RP3 (repaired) compared with Specimen
Next, from the ultimate moment capacity, the ultimate load P2 (control) demonstrate that external post-tensioned CFRP
Pu can be found from the following equation rods are able to compensate for partial or complete removal
of a prestressing strand.
4 Mu Although the repaired Specimen RP1 failed as a result
Pu = (9) of rupture of the external post-tensioned CFRP rods, this
(L − s) rupture occurred at deflections much greater than those
expected from service loads. Additionally, residual capacity
where s is the spacing between load points, 30 in. (762 mm), was present after CFRP rod rupture. This is significant in
as shown in Fig. 3. that catastrophic beam failure did not occur even though the
From Eq. (9), the theoretical ultimate capacity is 79 kip CFRP rods failed in tension. The repaired beams were essen-
(351 kN) for Specimen RP1 and 70 kip (311 kN) for Spec- tially as ductile as the control beam, because the mild steel
imen RP3 when implementing the predicted CFRP tendon and the remaining prestressing steel dominated the ductility
stresses at ultimate from the use of Eq. (3). Consequently, of the damaged beams.
the corresponding ratios of actual to theoretical ultimate load It was found that theoretical expressions from the litera-
are 1.42 and 1.46 for Specimens RP1 and RP3, respectively. ture may be used to predict the stress at ultimate in the CFRP
Similar to the ratio of 1.42 found for control Specimen P2, tendons used in this research as well as the ultimate capacity
these ratios show that theoretical ultimate loads are in good of the beams, with Eq. (3) being the preferred strain reduc-
agreement with actual measured ultimate loads, and that tion factor even though it was originally developed for steel
the design is conservative. A summary of the experimental tendons. Post-tensioning CFRP rods using unibody clamp
and theoretical ultimate loads is given in Table 3. Further- anchors and a mechanical stressing device is a viable tech-
more, the ratios of actual to theoretical ultimate load and the nique for the repair of concrete beams with severe damage
percentage of error between the actual and theoretical stress to internal steel prestressing tendons. It is recommended that
in the CFRP rods at ultimate indicate that Eq. (2) and (3) further studies be carried out to assess the system and its
are appropriate for predicting the stress in the CFRP rods at feasibility for general use.
ultimate when calculating the theoretical ultimate capacity It is also recommended that further studies be carried out
of prestressed concrete members repaired with the system to test unibody clamp anchors made of stainless steel and to
of unibody clamp anchors, mechanical stressing device, and determine the suitability of unibody clamp anchors for use
CFRP rods used in the current research. as a coupling device for CFRP rods. The successful imple-
mentation of the anchors and CFRP rods in this research
CONCLUSIONS suggests that the anchors could potentially be used to join
Based on the experiments carried out in this research, two sections of CFRP rod, facilitating post-tensioning of
it can be concluded that Specimens RP1 and RP3 were longer spans that are typical of actual bridges. Although the
successfully repaired using an external post-tensioning repair system investigated in this paper was successful, it
system consisting of CFRP rods, unibody clamp anchors, requires access to the end of the beam, which is not always
and a mechanical stressing device. Repaired Specimens RP1 available in field applications. For use of the CFRP repair

394 ACI Structural Journal/March-April 2014


system implemented in this research, however, only 18 in. precast and prestressed concrete buildings and bridges, and the application
of fiber-reinforced polymer composites.
(0.46 m) of free space is required behind the beams.
Lawrence D. Reaveley is a Professor and former Department Chair of
ACKNOWLEDGMENTS Civil and Environmental Engineering at the University of Utah. He received
The authors wish to acknowledge the financial support of the Utah his bachelor’s and MS degrees from the University of Utah, and his PhD
Department of Transportation and the University of Utah. The authors from the University of New Mexico, Albuquerque, NM. His research inter-
would also like to acknowledge the contributions of Hanson Structural ests include structural dynamics, with an emphasis on earthquake engi-
Precast and Sika, Inc. In addition, the authors would like to thank M. Bryant neering and seismic rehabilitation.
and R. Liu for their assistance in specimen fabrication and testing.
REFERENCES
AASHTO, 2009, “AASHTO LRFD Bridge Design Specifications,”
NOTATION fourth edition, American Association of State Highway and Transportation
ACFRP = area of post-tensioned CFRP rods
Officials, Washington, DC, 1518 pp.
ACFRP int = area of internal post-tensioned CFRP rods
Abdel Aziz, M.; Abdel-Sayed, G.; Ghrib, F.; Grace, N.; and Madugula,
ACFRP tot = total area of post-tensioned CFRP rods
M., 2005, “Analysis of Concrete Beams Prestressed and Post-Tensioned
Aps = area of internal steel prestressing strands
with Externally Unbonded Carbon Fiber Reinforced Polymer Tendons,”
A s = area of tensile mild steel reinforcement
Canadian Journal of Civil Engineering, V. 31, pp. 1138-1151.
a = depth of equivalent compression block
Abdelrahman, A. A.; Tadros, G.; and Rizkalla, S. H., 1995, “Test Model
cu = depth to neutral axis at ultimate
for the First Canadian Smart Highway Bridge,” ACI Structural Journal,
d = depth of mild steel reinforcement
V. 92, No. 4, July-Aug., pp. 451-458.
dCFRP = depth to CFRP rods
ACI Committee 440, 2004, “Prestressing Concrete Structures with FRP
de_CFRP = effective depth to CFRP rods
Tendons (ACI 440.4R-04) (Reapproved 2011),” American Concrete Insti-
ECFRP = modulus of elasticity of CFRP rods
tute, Farmington Hills, MI, 35 pp.
fc′ = compressive strength of concrete
ACI Committee 440, 2012, “Guide Test Methods for Fiber-Reinforced
fp_CFRP = stress in CFRP rods at ultimate
Polymer (FRP) Composites for Reinforcing or Strengthening Concrete and
fpe_CFRP = effective prestress in CFRP rods
Masonry Structures (ACI 440.3R-12),” American Concrete Institute, Farm-
fps = prestressing force in internal steel strands
ington Hills, MI, 23 pp.
fy = yield stress of mild steel reinforcement
Aravinthan, T.; Mutsuyoshi, H.; Fujioka, A.; and Hishiki, Y., 1997,
L = unbraced length of beam
“Prediction of the Ultimate Flexural Strength of Externally Prestressed PC
Mu = ultimate moment capacity
Beams,” Transactions of the Japan Concrete Institute, V. 19, pp. 225-230.
Rd = depth reduction factor
Burningham, C., 2011, “Development of a Carbon Fiber Reinforced
Sd = deviator spacing
Polymer Prestressing System for Structural Applications,” PhD disserta-
s = spacing between load points on top of beam
tion, University of Utah, Salt Lake City, UT, 103 pp.
β1 = factor relating depth of equivalent compression block to depth
El-Hacha, R., and Elbadry, M., 2006, “Strengthening Concrete Beams
of neutral axis
with Externally Prestressed Carbon Fiber Composite Cables: Experimental
Δ1 = displacement of uncracked section at load equal to ultimate load
Investigation,” PTI Journal, V. 4, No. 2, pp. 53-70.
Δ2 = displacement at failure
Elrefai, A.; West, J.; and Soudki, K., 2007, “Strengthening of RC Beams
εcu = failure strain of concrete in compression
with External Post-Tensioned CFRP Tendons,” Case Histories and Use of
Ωu = strain reduction factor
FRP for Prestressing Applications, SP-245, R. El-Hacha and S. H. Rizkalla,
eds., American Concrete Institute, Farmington Hills, MI, pp. 123-142.
AUTHOR BIOS Naaman, A., and Alkhairi, F., 1991, “Stress at Ultimate in Unbonded
Clayton A. Burningham is a PhD Candidate in the Department of Civil Post-Tensioning Tendons: Part 2—Proposed Methodology,” ACI Struc-
and Environmental Engineering at the University of Utah, Salt Lake City, tural Journal, V. 88, No. 6, Nov.-Dec., pp. 683-692.
UT. He received his bachelor’s and MS degrees from the University of Naaman, A.; Burns, N.; French, C.; Gamble, W.; and Mattock, A., 2002,
Utah. His research interests include repair of reinforced and prestressed “Stresses in Unbonded Prestressing Tendons at Ultimate: Recommenda-
concrete structures and post-tensioning carbon fiber-reinforced polymer tion,” ACI Structural Journal, V. 99, No. 4, July-Aug., pp. 518-529.
materials. Täljsten, B., and Nordin, H., 2007, “Concrete Beams Strengthened with
External Prestressing Using External Tendons and Near-Surface-Mounted
ACI member Chris P. Pantelides is a Professor in the Civil and Environ- Reinforcement,” Case Histories and Use of FRP for Prestressing Appli-
mental Engineering Department at the University of Utah. His research cations, ACI SP-245, R. El-Hacha and S. H. Rizkalla, eds., American
interests include seismic design and rehabilitation of reinforced concrete, Concrete Institute, Farmington Hills, MI, pp. 143-164.

ACI Structural Journal/March-April 2014 395


NOTES:

396 ACI Structural Journal/March-April 2014


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 111-S35

Study of Composite Behavior of Reinforcement and


Concrete in Tension
by John P. Forth and Andrew W. Beeby
This paper aims to further the understanding of the interaction
between reinforcement in tension and the surrounding cracked
concrete. This is achieved using the elastic analysis of axisymmetric
prisms reinforced with a single central bar. As a preliminary to the
analyses, the behavior of axially reinforced prisms is described
based on previous experiments. This preliminary analysis confirms
that the elastic analysis adopted in this investigation is reasonable.
Two analytical exercises are described: the first assumes no slip,
plasticity, or internal cracking at the interface between the steel
and the concrete; the second introduces internal cracking and
debonding between ribs. The first analysis indicates that shear
deformation of the surrounding concrete accounts for a substantial
proportion of the surface crack width, and therefore that this form
of deformation cannot be ignored in crack prediction formulae. The Fig. 1—Load-deformation response of Specimen 10T12
second analytical exercise shows that the internal cracking model (figure taken from Reference 4).
described by Goto is appropriate.
members quite effectively. The simpler model proposed
Keywords: axisymmetric tension specimens; cover; crack width calcula-
herein further confirms the concept that crack width is a
tions; cracking mechanisms; finite element modeling; shear distortions.
function of the shear deformation of the concrete cover.
INTRODUCTION
BASIC BEHAVIOR AS REVEALED BY TESTS
The objective of this study is to gain greater under-
The information used herein is taken from Reference 4.
standing of the interaction of reinforcement and concrete in
Initially, strain data for Specimen 100T12 will be presented,
tension. The analysis models used have been kept as simple
as this gives a convenient illustration of a number of aspects
as possible; the approach has been limited to pure elastic
of behavior. Figure 1 shows the load-average reinforcement
behavior, and an assumed internal cracking pattern, based
strain response for this specimen.
on the work by Goto,1 has been adopted (rather than use
It should be noted that the response is not a continuous
a nonlinear finite analysis software based on, for instance,
smooth curve as is commonly plotted, but is made up of a
a smeared cracking approach where cracks are predicted
series of linear segments separated by a sudden increase in
regions of damaged material with degraded properties). This
strain on the occurrence of each crack. Up to a load of approx-
keeps difficulties in interpretation to a minimum, though it
imately 7.868 kip (35 kN), these linear segments, extrapo-
is recognized that concrete does not necessarily behave in a
lated backwards, can be seen to pass through the origin. The
perfectly elastic manner.
behavior of the tension specimen with a given number of
Use of elastic modeling, where the cracks being studied are
cracks is thus elastic. Using the computer to produce best
open, is not so unreasonable as might be thought by some.
fit lines for each segment enables the stiffness of the spec-
Extensive data obtained by Scott and Gill2 and Beeby and
imen to be established for each crack configuration. Figure 2
Scott3-5 suggest that much of the behavior revealed during
shows this compliance plotted against the number of cracks.
tension tests is close to what would be expected from an
There is a linear relationship between stiffness and number
assumption of elastic-brittle behavior for concrete in tension.
of cracks. This implies that the formation of each crack
This is effectively what will be assumed in the study.
reduces the stiffness of the element by a constant amount.
The final point for four cracks does not quite fit the linear
RESEARCH SIGNIFICANCE
relationship. This point is obtained from the behavior imme-
In reinforced concrete, the interaction between reinforce-
diately after formation of the fourth crack. Figure 1 shows
ment in tension and the surrounding concrete is still not
that, at higher loads, there are two further sudden increases
fully understood. Two internal failure mechanisms, pure slip
in strain. These increases were not related to the formation
and internal cracking, form the basis of three approaches
that exist in the codes to model the tension zone behavior ACI Structural Journal, V. 111, No. 2, March-April 2014.
under service loads. The models presented in this investiga- MS No. S-2012-148, doi:10.14359.51686564, was received May 10, 2012, and
reviewed under Institute publication policies. Copyright © 2014, American Concrete
tion are based only on Goto’s internal cracking mechanism. Institute. All rights reserved, including the making of copies unless permission is
obtained from the copyright proprietors. Pertinent discussion including author’s
These models predict the experimental behavior of tension closure, if any, will be published ten months from this journal’s date if the discussion
is received within four months of the paper’s print publication.

ACI Structural Journal/March-April 2014 397


Fig. 2—Stiffness of element as function of number of cracks
for Specimen 100T12.

Fig. 4—Concrete stresses at right end of Specimen T16B1.


(Note: 1 MPa = 145 psi; 1 kN = 0.225 kip; 1 mm = 0.0394 in.)

Fig. 3—Specimen T16B1: reinforcement strains along bar at


various loads. (Note: 1 kN = 0.225 kip; 1 mm = 0.0394 in.)
of visible surface cracks, and it may be speculated that they
arise from some form of internal failure. It should be noted Fig. 5—Relationship between So and cover. (Note: 1 mm =
that both these increases in strain occurred only at very high 0.0394 in.)
levels of stress in the reinforcement (>58.02 ksi [400 MPa]).
Figures 1 and 2 show that the behavior between cracking A critical factor in crack prediction theories is the defini-
events is generally elastic, and this is the assumption that tion of the transfer length. This is the distance on one side of
will effectively be made in the finite element analyses. a crack over which the stress in the reinforcement is affected
The next aspect of behavior to be considered is the vari- by the crack. In References 3 through 5, this is represented
ation of steel stress or concrete stress with distance from a by the symbol So. In some papers, the symbol ltr is used.
crack. Figure 3 shows the variation of strain along a rein- There are various ways of estimating So from experimental
forcing bar for various levels of axial load. Various pieces of results. Most of these are indirect and assume a relation-
information can be gleaned from this figure. ship, for example, between So and crack width. The work
First, over a considerable part of the distance from a crack reported in References 3 through 5, however, recorded the
or the free end of the specimen, the variation in strain is very variation in strain at closely spaced intervals along the rein-
close to linear. This can possibly be seen better from Fig. 4, forcing bar, permitting the direct measurement of So. Even
which shows the variation in concrete stress over the end with these tests there are difficulties, as can be seen from the
11.81 in. (300 mm), enlarged for two levels of load. At the strain variation for the 4.496 kip (20 kN) level of load shown
lower load level, there is a clear curve over the part of the in Fig. 4. It is difficult to define exactly the point where the
bar where the stress is close to that for uncracked concrete. crack no longer influences the strain. The method used to
This is not clear for the higher load, where it would not be establish a value for So is illustrated in Fig. 4, where So is
unreasonable to consider the relationship to be linear over defined as the distance from the crack (or free end of the bar)
the whole distance to midway between cracks. and the point where a best fit line through the strains inter-
Secondly, even at the low level of load, which is below sects the strain in the uncracked concrete. This is clearly an
the cracking load for the specimen, the strain in the rein- imperfect procedure, but it is consistent and seems to agree
forcement over most of the length affected by the end of well with the calculation of So by indirect means. A good,
the specimen So is considerably greater than the cracking straight line relationship was found between So and cover.
strain of the concrete, which can be assumed to be in the This is illustrated in Fig. 5, which also includes the equation
range of 100 to 150 × 10–6. This implies that some form of for the straight line.
internal failure occurs over the whole length So as soon as An issue that has been studied by a number of researchers,
cracking occurs. but is generally ignored by those developing theories of

398 ACI Structural Journal/March-April 2014


Fig. 7—Variation in calculated stresses in concrete with
Fig. 6—Schematic illustration of situation analyzed. distance from free end. (Note: 1 MPa = 145 psi; 1 mm =
0.0394 in.; S*0s is surface spacing; S*0m is mean spacing.)
cracking, is the shape of the crack (that is, how the width
of a crack varies between the bar surface and the surface scx = (T – Asssx)/Ac (1)
of the concrete). An assumption generally seems implicit
in cracking formulae based on the classical theory that The average stress in the concrete was calculated in this
the crack width at the concrete surface is the same as that way because it corresponds to the method of concrete stress
at the bar surface. There is now ample evidence that this calculation used in References 3 through 5. It is also effec-
is not the case. The research evidence has been reviewed tively what is used in many theoretical derivations of crack
in Reference 6, and this shows that the cracks are tapered, prediction formulas.
being much smaller near the bar surface than at the concrete There are several interesting points which arise from
surface. Though the results are highly variable, it can be Fig. 7. First, as might be expected, the surface stress is not
concluded that the width at the concrete surface is at least the same as the average stress, but is considerably lower over
twice that near the bar surface. almost the whole of the length analyzed. So for the surface is
thus different than So for the average stress, with the surface
SIMPLE ELASTIC ANALYSIS WITHOUT value being considerably longer. Straight lines have been
INTERNAL CRACKING drawn in passing through the origin and the point where
Initially, a simple two-dimensional analysis was performed the calculated curves reach two-thirds of the homogeneous
on an axially reinforced circular cross section prism to give stress. This is simply done to permit a simple visual compar-
some idea of the stress and deformation conditions around a ison of the curves. It should also be noted that the surface is
bar. The axial symmetry greatly simplified the analysis. A actually in compression for the area closest to the crack face.
fully elastic analysis of the area surrounding a bar, assuming Second, the stress in the concrete does not actually reach the
no slip between the bar and the concrete, was considered. stress calculated for a homogeneous section. Thus, there is
The situation analyzed is illustrated in Fig. 6, in which ws is no absolutely clear definition of So, as is assumed in all theo-
the surface crack width. retical equations for predicting cracking. This is not neces-
Analyses were carried out using axisymmetric elements. sarily a critical point, but it may be worth remembering that
For the initial analysis, a 0.79 in. (20 mm) diameter f bar was So is an effective value rather than an absolute value.
considered with 1.97 in. (50 mm) cover c. Square elements Figure 8 shows the variation in the deformed shape of the
of 0.20 in. (5 mm) were used, and a length from the free free end over the height of the crack from the bar surface
face up to the fixed end of 5.91 in. (150 mm) was assumed. to the specimen surface. Quantitative comparisons cannot
A uniform stress of 14.51 kip/in.2 (100 MPa) was applied to directly be made in this case, as the geometries of the avail-
the free end of the bar. Figure 7 shows the stress distribution able experimental specimens are somewhat different from
in the concrete along the specimen obtained in two ways: the that analyzed. The deformation at the surface in Fig. 8 corre-
stress in the concrete on the outer face (that is, the concrete sponds to a crack width of 0.0019 in. (0.047 mm).
surface), and the average stress in the concrete. The average Analyses have been carried out for different covers,
stress was calculated by taking the force in the reinforcing and Fig. 9 shows the calculated crack widths as a function
bar at each 0.20 in. (5 mm) section, deducting this from the of cover.
total applied force at the free end, and dividing this differ- Crack width decreases with a decrease in cover. The
ence by the area of the concrete. In algebraic terms, if T is decrease is not linear as suggested from the experimental
the tension force applied at the free end, the bar area is As, data3-5; however, certain factors should be borne in mind.
and the concrete area is Ac, the stress in the bar at any section The finite element analyses are elastic, and therefore, any
a distance x from the crack is ssx and the average stress in specimen having a geometrically similar cross section to the
the concrete is scx then, by equilibrium, because the force at one for which the crack width has been calculated will give a
section x is T, the stress in the concrete is given by crack width that can be calculated by direct scaling from the
previously calculated width. Thus, for example, the crack

ACI Structural Journal/March-April 2014 399


Fig. 8—Variation in calculated crack width with distance Fig. 9—Variation in maximum surface crack width with
from bar surface. (Note: 1 mm = 0.0394 in.) cover. (Note: 1 MPa = 145 psi; 1 mm = 0.0394 in.)

width for any specimen with a value of c/f of 2.5 will lie on a
straight line joining the point for c/f = 2.5 to the origin. This
applies for any other value of c/f. Thus, all results for any
specimens with c/f in the range 1 to 2.5 will lie between the
two dashed lines drawn in Fig. 9. If a relatively random set
of tension specimens are analyzed, the result, when plotted
on a graph such as Fig. 9 will, to the engineering eye, be
accepted as giving a linear relationship between cover and
crack width with some relatively small level of scatter.
Figure 10 aims to make an approximate quantitative
comparison between calculated and experimental crack
widths. The test specimens, from Farra and Jacccoud,7 were
3.94 in. (100 mm) square and reinforced with a single axial
Fig. 10—Maximum crack widths from Farra and Jaccoud
0.79 in. (20 mm) bar. The cover was thus 1.57 in. (40 mm),
Specimens N-20-207 compared with finite element analysis.
and results from an analysis for 1.57 in. (40 mm) cover have
(Note: 1 MPa = 145 psi; 1 mm = 0.0394 in.)
been used in the comparison. It should be remembered,
however, that the experimental specimens had a square but will undergo some plastic deformation before rupture.
cross section, whereas this analysis considered a circular This will result in the actual deformation of the concrete
cross section. It can be seen that the analysis underestimates being greater than that calculated on the assumption of elas-
the maximum crack width by approximately 30%. This is to ticity. Additionally, a short-term value has been used for Ec.
be expected, as no account has been taken in this analysis There is likely to be some creep during the test that would
of internal failure (slip or internal cracking) which, as has result in further deformation of the concrete and steel, and
been discussed previously, occurs and reduces the stiffness hence higher calculated crack widths. Depending on the
of the concrete in tension. This will be considered further in effect of these two factors, the calculated width could be
a following section. closer to the experimental values.
It seems likely that this initial simple analysis gives a Overall, the analysis seems to have been very successful
lower-bound indication of the deformation of the tensile in predicting the general qualitative behavior of axially rein-
concrete, and hence, the estimate of the crack width. In forced specimens.
reality, concrete in tension is not absolutely elastic-brittle,

400 ACI Structural Journal/March-April 2014


Fig. 11—Finite element model of axially reinforced tension
specimen with internal cracks.
ANALYSIS OF TENSION ZONE INCLUDING
INTERNAL CRACKS
As mentioned previously, some form of internal failure
takes place over the whole of the length So from a very early
load stage. There are two mechanisms that are commonly
proposed for this internal failure: slip along the steel-con-
crete interface, and internal cracks forming at an angle to the Fig. 12—Axisymmetric model created using OASYS-GSA
axis of the bar. Slip is the most common mechanism invoked, software.
and has been used as the basis for many crack prediction
theories. The internal cracking mechanism was first illus-
trated by Goto,1 and was elaborated further by Beeby and
Scott,4 whose model which will be investigated in this paper
(it is believed that there are plenty of advocates of the slip
model who can, if they wish, carry out further modeling of
this option). It should be noted that the type of modeling
that will be attempted herein will not prove that a particular
model is the actual behavior; at best, it can simply show that
the particular model gives a reasonable simulation of reality.
Other models may exist that are as good or better. It could,
however, show that a particular model was unreasonable.
An axially symmetrical specimen was chosen for the
analysis. Initially, the analysis will be carried out on a spec-
imen the same basic size as that used to produce the results
detailed in Fig. 7 and 8. It will now, however, also model a
number of internal cracks. The length of the model specimen Fig. 13—Axisymmetric model: detail of assumed crack
has been doubled to 11.81 in. (300 mm), partly because the pattern.
length So was expected to be greater than in the case with
no internal cracks, and partly because it was felt that the cracking (for example, Gerstle and Ingraffea9). The differ-
length used in the previous analysis may have been slightly ence between those analyses and this analysis is that most
short for the largest cover. The number of cracks, the angle of the other analyses have attempted to study the develop-
of the cracks to the bar axis, and the length of the cracks ment of the internal cracking as a function of applied load,
is somewhat arbitrary, but is based on photographs from whereas in this study, to study a larger range of variables, a
Goto1 and Otsuka and Ozaka,8 and the analysis presented crack pattern has been assumed.
in Beeby and Scott.4 Experimental work by Goto and others The results from the analysis are shown in Fig. 13 to 18.
suggest that internal cracks form at each rib on the bar. The Figure 15 shows a number of interesting changes from the
spacing used in the model analyzed herein is rather larger stress results shown in Fig. 7 resulting from the element
than the rib spacing for reasons of practicality. An angle of without internal cracks. First, the relationship between the
60 degrees to the axis of the bar was chosen, although this average stress and distance from the crack is much closer to
angle could only be approximately maintained as the basic linear. It now models more closely the experimental result
grid of 0.20 in. (5 mm) did not permit exactly 60 degrees shown in Fig. 4.
to be maintained for all cracks. Furthermore, the aim was The deformed shape of the free end (Fig. 16), which is
to use a linear decrease in crack height with distance from actually a tracing of the finite element analysis graphical
the crack face. Again, the 0.20 in. (5 mm) grid meant that output (Fig. 14) with the elements and nodes removed for
this could only be achieved approximately. The elastic clarity, is now possibly less similar to that obtained for the
finite element model used is illustrated in Fig. 11 and 12. specimen without internal cracks and to that of the exper-
Others have carried out analyses aimed at studying internal imental specimen, but it is still reasonable. The surface

ACI Structural Journal/March-April 2014 401


Fig. 16—Calculated deformation of specimen with seven
internal cracks. (Note: Dimensions in mm; 1 mm =
0.0394 in.)
Fig. 14—Axisymmetric model: deformed state. (Note:
1 MPa = 145 psi.)

Fig. 17—Maximum crack widths from Farra and Jaccoud


Specimen N-20-207 compared with finite element analysis,
Fig. 15—Concrete stresses calculated by finite element with and without internal cracks. (Note: 1 mm = 0.0394 in.;
analysis for specimen with seven internal cracks. (Note: 1 MPa = 145 psi.)
1 MPa = 145 psi; 1 mm = 0.0394 in.)
deformation corresponds to a crack width of 0.0031 in.
(0.08 mm), 60% greater than that obtained for the specimen
without internal cracks. This agrees closely with the experi-
mental results, as can be seen from Fig. 17, which shows the
same data as used for Fig. 10, but with the calculated line
shown for the analysis with internal cracks.
The predicted effect of cover is shown in Fig. 18, where it
can be seen that a reasonably linear relationship is predicted
between cover and surface crack width. There is, however,
some scatter in these results, possibly due to the difficulty of
modeling absolutely geometrically similar internal cracks in
the analyses for the various covers.
A further assessment of the performance of the model Fig. 18—Predicted maximum crack width as function of
can be performed by considering the work by Broms10 and cover for analyses including internal cracks. (Note: 1 mm
Beeby.11 Broms10 carried out a series of tests on short prisms = 0.0394 in.)
and measured the longitudinal extension at various stress
levels in the reinforcement. Results are presented in Refer- and calculated results for two levels of stress. The exper-
ence 10 for a circular cross section specimen, 6 in. (152 mm) imental results have been scaled from a figure in Broms’
in diameter and 8 in. (203 mm) long with a central 1 in. paper.10 It can be seen that, in this case, the experimental
(25 mm) diameter bar. An elastic analysis has been carried results exceed the calculated results by approximately 20%.
out for this specimen, and Fig. 19 shows the experimental The general trend of the results, however, is well reflected by
the calculations. This is not in absolute agreement, but it is

402 ACI Structural Journal/March-April 2014


Fig. 19—Comparison of calculated and measured crack Fig. 20—Comparison of calculated and measured overall
widths for Brom’s Specimen T-C-5.10 (Note: 1 mm = extension for specimen reported in Reference 11. (Note:
0.0394 in.; 1 MPa = 145 psi.) 1 mm = 0.0394 in.; 1 MPa = 145 psi.)
probably within the range that could be covered by judicious tions implicitly assume that this shear deformation is negli-
adjustment of the model. gible. The analyses show that this is not so; the elastic shear
A cylindrical specimen was tested by Beeby11 where the deformation of the concrete in the analyses reported herein
overall extension of a 5.91 in. (150 mm) diameter specimen accounts for around two thirds of the crack width. Had other
with an axial 0.87 in. (22 mm) bar at various distances from material factors, such as creep or inelasticity of the concrete
the bar surface were measured. This has been analysed, and in tension, been taken into account in the analyses, the shear
the extensions scaled off the figure presented in Reference deformations and their contribution to the crack widths
11. Figure 20 shows the measured extensions compared with would have been even greater. This substantial contribution
the finite element calculations for two levels of steel stress. of the shear deformation of the cover concrete seems ines-
Agreement between experiment and calculation is slightly capable, and suggests that any approach to the prediction of
better than for Broms10 in Fig. 19, though the calculation, crack widths that ignore this are fundamentally flawed.
again, tends to underestimate the measured results. In the second analysis, the output from the model with
The introduction of the internal cracks (Fig. 12 to 14) internal cracks generally agreed with the experimental
has, in general, resulted in an improved agreement with the data, where comparisons could be made. This suggests
experimental behavior, and does suggest that the internal that the internal cracking model of behavior can provide
cracking model is a reasonable model for the behavior of a good model of cracking behavior. It does not prove that
tension zones. The analysis is clearly capable of further the mechanism accommodating excess tensile strains above
refinement, and has only been carried far enough to demonstrate those which the concrete can support in tension is internal
its inherent reasonableness. cracking; it shows that it is a viable alternative to the bond-
slip model, and should not be dismissed.
DISCUSSION The analyses carried out are somewhat limited, and can be
considered to make a prima facie case for the reasonableness
Elastic analyses of the internal cracking model rather than a fully developed
Two basic analyses have been described in this paper. analytical study. Some of the more obvious limitations of the
In the first, the concrete is considered to remain elastic model are given as follows.
and uncracked, and complete bond is assumed between the • Circular cross sections are analysed, not square or rect-
reinforcement and the concrete. In the second, a pattern of angular ones. Due to difficulties in manufacture, very
internal cracking has been assumed, based on the findings few circular specimens have been made and tested;
of Goto.1 In neither of the analyses is any form of bond-slip thus, it is not possible to compare the analytical results
relationship assumed; thus, bond-slip can have no influence rigorously with test results that are almost all from spec-
on the results obtained. imens with square or rectangular cross sections.
In the first analysis (without internal cracking), it was • Location and size of internal cracks is somewhat arbi-
expected that the predicted crack widths would be less than trary. As mentioned in a previous section, no attempt has
obtained experimentally, and this proved to be the case. been made to refine the form of the internal cracking.
Nevertheless, the analyses were not trivial, and the results From the existing experimental evidence, the pattern
illustrate a significant point that has commonly been ignored. assumed seems reasonable, but it cannot be said to be
If a shear stress is applied to a material, then shear strains rigorously justified.
and displacements occur. Bond stress is simply a shear • Rib pattern. By its nature, the axisymmetric analysis
stress, and therefore, the concrete surrounding a bar in the assumes that the ribs are perpendicular to the bar axis
region of a crack undergoes shear deformations. Though this and extend round the full circumference of the bar.
has not been shown to be explicitly stated, the classical theo- This is not normally so for modern ribbed bars, where
ries of cracking that lie behind many crack prediction equa- the ribs tend to be staggered. There is also frequently

ACI Structural Journal/March-April 2014 403


Table 1—Comparison of crack widths at center analysis of three-dimensional specimens (with the exception
and near corner of axially reinforced tension of axisymmetric situations).
specimens (from Reference 12)
Issues relating to development of valid design
formulae for crack width prediction
It would be helpful for further discussion if a brief
outline is given of the development of crack theories and
code provisions.
The earliest developed theory of cracking assumed that
Mean values of w/e 5% values of w/e the widths of cracks accommodated slip between the bar
Specimen B a b b/a a b b/a and the concrete. The theory ignored any contribution to
the crack width from the shear deformation of the cover
Z2 80 76 113 1.48 174 203 1.17
concrete. To develop equations, it required assumptions be
Z6 130 104 171 1.64 253 354 1.40 made about the development of the bond stresses as a func-
Z7 180 91 254 2.79 231 535 2.32 tion of slip. Many different assumptions were considered,
Z9 230 101 259 2.56 282 580 2.06
but all resulted in a basic equation of the form

Notes: B, a, and b, are in mm; 1 mm = 0.0394 in.; w/e is average crack width/average w = kffcte/tr (2)
surface strain.

a longitudinal rib. The result of this is that, in reality, where w is crack width (variously defined); k is a constant;
the pattern of internal cracks may be considerably more fct is the tensile strength of the concrete; e is strain (variously
complicated than is modeled in this analysis. defined); f is the bar diameter; r is reinforcement ratio (vari-
• The spacing of the internal cracks, which would be ously defined); and t is bond strength.
expected to follow the spacing of the ribs, is too large Many design provisions have been based directly on
in the model. this equation, including those in the CEB-FIP Model
Code 1990.18
Issues requiring further study In 1965, Broms19 and Broms and Lutz20 published a radi-
What is measured when crack widths are being investi- cally different theory which assumed that the crack width
gated is the crack width on the surface. It is found that the arose entirely from the shear deformation of the cover
surface crack width is strongly dependent on where the concrete. The following formula was developed
cracks are measured relative to the position of the reinforce-
ment. If the cracks are measured at points on the surface wav = 2tes (3)
directly over the bar, they are found to be substantially
smaller than if they were measured, for example, close to or
the corner of an axially reinforced prism. The variation is
less in situations where there are multiple bars and the crack
width over the bars is compared with that at mid-spacing. wmax = 4tes
This behavior is illustrated in Table 1, containing data from
Reference 12. where wav is the average crack width; wmax is the maximum
This effect seems perfectly rational for crack widths crack width; t is distance from the center of the bar to the
resulting from shear deformation of the concrete; the shear point on the surface where the crack width is considered; and
displacement will increase with increasing distance from the es is average strain of the steel.
bar in any direction. Because the corner of the specimens For multiple bars, t was modified to te, an effective distance,
used in Table 1 are further from the bar than a point directly which is defined, for bottom cracks, as 3 (tb A), where tb is the
over the bar, the deformation will be greater, and the crack distance from the bottom of the beam to the center of the
width larger. This effect was recognized by Broms10 and in lowest layer of bars, and A is the area of concrete immedi-
the work carried out at the Cement and Concrete Associa- ately surrounding the tension reinforcement. This formula,
tion.12,13 It is implicitly included in the ACI code14 formula along with many others, was tested against the available
and taken into account directly in the UK code.15 crack width data by Gergely and Lutz21 and shown to be the
The analyses performed herein have been exclusively best available at the time. The formula has formed the basis
concerned with members subjected to pure tension. There of the ACI code14 crack width control provisions ever since.
is evidence that flexural members behave rather differently. At the same time as the work being carried out at by
Studies by Beeby16 showed that, in shallow members, such Broms, Lutz and Gergely,19-21 a major series of tests were
as slabs, the depth of the tension zone has a significant effect carried out at the Cement and Concrete Association in the
on the crack width. This is explicitly taken into account in UK.13 The first series of tests consisted of 105 beams, and
the UK code,15 and is also recognized in Eurocode 2.17 the results were with the publishers at the time that Broms’19
These effects have not been investigated in this study paper appeared. The paper13 concluded that crack width
because the finite element package used did not permit the could be predicted by the formula

404 ACI Structural Journal/March-April 2014


wmax = 3.3acrem (4) in formulae derived on this basis. Strains beyond those
supportable by concrete in tension are accommodated by a
where wmax is the maximum crack width; acr is the distance reduction in the stiffness of the cover concrete by internal
from the point where the crack width is being considered to cracking. This is the case for the UK15 and ACI codes14 and
the surface of the nearest bar; and em is the average strain at the codes of any country that basically follow either British
the level where the width is considered. or American practice; and
It can be seen that, while acr is slightly larger than t in 3. The assumption that the crack width arises from a
Broms’19 equation and that the equation aims to predict combination of slip and deformation of the concrete. This
the maximum width rather than the average, the Cement is true of CEB-FIB Model Code 197824 and Eurocode 2,17
and Concrete Association13 and Broms’19 basic formula and will become the case for all countries which either adopt
are almost the same. Like Broms’19 theoretical approach, Eurocode 2 or base their national codes on Eurocode 2.
the formula assumes that there is no bond failure or slip at From this investigation, it is apparent that a significant
the bar-concrete interface, and that the crack width results proportion of the crack width is due to the deformation of the
entirely from the deformation of the cover concrete. The concrete surrounding the bar. Therefore, Item 1 mentioned
Cement and Concrete Association13 work was extended over previously is not tenable as a basis for a rational crack
the following few years by Beeby,22-24 resulting in modifica- prediction formula; Items 2 and 3 are tenable depending on
tions to deal with bar spacing and the effect of the depth of whether or not slip plays a significant role. The paper does
the tension zone. It was recognized that some mechanism not aim to show which of these possibilities is closest to the
was necessary to accommodate the strains in the concrete truth, merely that the concept of crack width being a func-
in excess of the tensile strain capacity of the concrete, and it tion of the deformation of the cover concrete is reasonable.
was proposed that the form of cracking identified by Goto1
provided that mechanism. The resulting formula was simpli- CONCLUSIONS
fied somewhat, and has been used in UK codes15 since 1972. In this paper, a number of simple elastic finite element
This formula is analyses of the concrete in tension surrounding tensile rein-
forcement, in members subject to pure tension, are described
w = 3acrem/{1 + 2(acr – c)/(h – x)} (5) and the results compared with the behavior of actual tension
specimens, as revealed by experiment. The study leads to the
where c is the cover to the face of the member where the following conclusions.
crack width is being considered; h is the overall depth of 1. The analyses show clearly that cover should be an
the section; and x is the neutral axis depth (depth of the important factor in any approach to the calculation of crack
compression zone). widths. This effect arises from the shear distortion of the
During the same time, Ferry-Borges25 developed a concrete between the bar surface and the concrete surface.
formula that combines the theoretical ideas behind the clas- 2. Experimental results show clearly that there should be
sical bond-slip model and the shear deformation models. His some form of internal failure in the region of the bar over
formula is the whole length over which the crack influences the stress
distribution. Two mechanisms have been proposed for this
sav = k1c + k2f/r (6) internal failure: slip along the bar-concrete interface, and
internal cracks of the form proposed by Goto1 and elabo-
wav = savem rated by Beeby and Scott.4 In this paper, Beeby and Scott’s
model is analyzed and is shown to describe the experimental
where sav is the average crack spacing. behavior of tension members effectively.
The term k1c is justified in Reference 25 by the following 3. The results from this study and those described in
statement: “The need to consider the influence of the thick- References 2 through 5 suggest the possibility of a model for
ness of the cover, c, is easy to understand. In fact, even if tension zone behavior under service loads which is, in prin-
perfect bonding between concrete and steel existed, the ciple, simpler and more all-embracing than current models.
mean distance between cracks would not be zero but propor- This can be described by the following two assumptions:
tional to c.” a) Concrete in tension behaves in an elastic-brittle
This formula was adopted in the CEB Model Code 1978,26 manner; and
and also in Eurocode 2.17 b) Force is transferred between ribbed reinforcing bars
The object of this survey of approaches to crack width and concrete by the mechanical action of the ribs. It is
calculation is to point out that there are three basic approaches assumed that there is no bond between the ribs and no slip
used in codes: past the ribs.
1. The assumption that the crack width arises purely from
slip. This is used in a number of codes and, notably, in AUTHOR BIOS
CEB-FIB Model Code 199018; John P. Forth is a Senior Lecturer in the School of Civil Engineering
at the University of Leeds, Leeds, UK. He received his BEng in civil and
2. The assumption that there is no significant slip and structural engineering from the University of Sheffield, Sheffield, UK. He
that the crack width arises entirely from the deformation of received his PhD from the University of Leeds. His research interests
the concrete around the bar. Bond and slip do not feature include serviceability and durability performance of reinforced concrete
and masonry structures.

ACI Structural Journal/March-April 2014 405


The late Andrew W. Beeby was Emeritus Professor of Structural 13. Base, G. D.; Beeby, A. W.; Read, J. B.; and Taylor, H. P. J., “An
Design in the School of Civil Engineering at the University of Leeds. Investigation of the Crack Control Characteristics of Various Types of Bar
He received both his first degree and his PhD from London University, in Reinforced Concrete Beams,” Research Report 18, Cement and Concrete
London, UK. He research interests included the serviceability behavior of Association, London, UK, Dec. 1966, 31 pp.
concrete structures. 14. ACI Committee 318, “Building Code Requirements for Structural
Concrete (ACI 318-11) and Commentary,” American Concrete Institute,
REFERENCES Farmington Hills, MI, 2011, 503 pp.
1. Goto, Y., “Cracks Formed in Concrete around Deformed Tension 15. BS8110-2, “Structural Use of Concrete—Part 2: Code of Practice for
Bars,” ACI Journal, V. 68, No. 4, Apr. 1972, pp. 224-251. Special Circumstances,” British Standards Institution, BSI Milton Keynes,
2. Scott, R. H., and Gill, P. A. T., “Short Term Distributions of Strain London, UK, 1985, 62 pp.
and Bond Stress along Tension Reinforcement,” The Structural Engineer, 16. Beeby, A. W., “An Investigation of Cracking in Slabs Spanning One
V. 65B, No. 2, June 1987, pp. 29-43. Way,” Technical Report TRA 433, Cement and Concrete Association, Apr.
3. Beeby, A. W., and Scott, R. H., “Insights into the Cracking and 1970.
Tension Stiffening Behavior of Reinforced Concrete Tension Members 17. Eurocode 2, “Design of Concrete Structures – Part 1-1: General
Revealed by Computer Modeling,” Magazine of Concrete Research, V. 56, Rules and Rules for Buildings,” CEN, EN 1992-1-1, Brussels, Belgium,
No. 3, Apr. 2004, pp. 179-190. 2004, 225 pp.
4. Beeby, A. W., and Scott, R. H., “Cracking and Deformation of Axially 18. CEB-FIP, “CEB-FIP Model Code (1990),” Bulletin d’information
Reinforced Members Subjected to Pure Tension,” Magazine of Concrete No. 213/214, Comité Euro-Internationale du Béton (CEB), Lausanne, Swit-
Research, V. 57, No. 10, Dec. 2005, pp. 611-621. zerland, 1993, 437 pp.
5. Beeby, A. W., and Scott, R. H., “Mechanisms of Long-Term Decay 19. Broms, B. B., “Crack Width and Crack Spacing in Reinforced
of Tension Stiffening,” Magazine of Concrete Research, V. 58, No. 5, June Concrete Members,” ACI Journal, V. 62, No. 10, Oct. 1965, pp. 1237-1256.
2006, pp. 255-266. 20. Broms, B. B., and Lutz, L. A., “Effects of Arrangement of Reinforce-
6. Beeby, A. W., “The Influence of the Parameter f/reff on Crack ment on Crack Width and Spacing of Reinforced Concrete Members,” ACI
Widths,” Structural Concrete, V. 5, No. 2, 2004, pp. 71-83. Journal, V. 62, No. 11, Nov. 1965, pp. 1395-1410.
7. Farra, B., and Jaccoud, J.-P., “Influence du beton et de l”armature sur 21. Gergely, P., and Lutz, L. A., “Maximum Crack Width in Rein-
la fissuration des structures en beton,” Publication No. 140, Rapport des forced Concrete Flexural Members,” Causes, Mechanism, and Control of
essais de tirants sous deformation impose de court duree, Department de Cracking in Concrete, SP-20, R. E. Philleo, ed., American Concrete Insti-
Genie Civil, Ecole Polytechnique Federale de Lausanne, Nov. 1993. tute, Farmington Hills, MI, 1968, pp. 87-117.
8. Otsuka, K., and Ozaka, Y., “Group Effects on Anchorage Strength 22. Beeby, A. W., “An Investigation of Cracking in the Side Faces of
of Deformed Bars Embedded in Massive Concrete Block,” Proceedings of Beams,” Technical Report 42.466, Cement and Concrete Association, Dec.
International Conference on Bond in Concrete—From Research to Prac- 1971, 11 pp.
tice, V. 1, Riga Technical University, Riga, Latvia, Oct. 1992, pp. 1.38-1.47. 23. Beeby, A. W., “A Study of Cracking in Reinforced Concrete
9. Gerstle, W., and Ingraffea, A. R., “Does Bond-Slip Exist?” Concrete Members Subjected to Pure Tension,” Technical Report 42.468, Cement
International, V. 13, No. 1, Jan. 1991, pp. 44-48. and Concrete Association, June 1972, 25 pp.
10. Broms, B., “Theory of the Calculation of Crack Width and Crack 24. Beeby, A. W., “The Prediction of Crack Widths in Hardened
Spacing in Reinforced Concrete Members,” Cement och Betong, No. 1, Concrete,” The Structural Engineer, V. 57A, No. 1. Jan. 1979, pp. 9-17.
1968, pp. 52-64. 25. Ferry-Borges, J., “Cracking and Deformability of Reinforced
11. Beeby, A. W., Concrete in the Oceans: Cracking and Corrosion, Concrete Beams,” V. 26, International Association for Bridge and Struc-
CIRIA/UEG, Cement and Concrete Association, Department of Energy, tural Engineering. Zurich, Switzerland, 1966, pp. 75-95.
1978, 77 pp. 26. “CEB-FIP Model Code for Concrete Structures,” Bulletin d’infor-
12. Beeby, A. W., “A Study of Cracking in Reinforced Concrete mation No. 125, Comité Euro-International du Beton, Paris, France, Apr.
Members Subjected to Pure Tension,” Technical Report No. 42.468, 1978, 460 pp.
Cement and Concrete Association, June 1972, 25 pp.

406 ACI Structural Journal/March-April 2014


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 111-S36

Flexural Capacity of Fiber-Reinforced Polymer


Strengthened Unbonded Post-Tensioned Members
by Fatima El Meski and Mohamed Harajli
Experimental and analytical investigations were carried out
 d − c
for evaluating the nominal moment capacity of unbonded post- e s = e cu  (6)
tensioned members when strengthened using external fiber-  c 
reinforced polymer (FRP) composites. In the experimental part of
the study, 36 simply supported specimens were tested to failure. The
main test parameters included area of internal tension reinforce-  d f −c 
ment, area of external FRP reinforcement, span-depth ratio of the e f = e cu  − e bi ≤ e fd (7)
member, profile of the prestressing tendons, and type of concrete  c 
structural system. In the analytical part of the study, a design-
oriented procedure for evaluating the nominal moment capacity
of FRP-strengthened post-tensioned members with internal or
fc′
external unbonded tendon systems is developed. The procedure e fd = 0.41 ≤ 0.9e fu (8)
is consistent with the approach proposed in ACI Committee 440 nf Ef t f
report for reinforced concrete or bonded prestressed concrete
members, and is applicable for both simply supported and contin-
uous members. The accuracy of the design approach was verified
Aps f ps + As fs + A f ( f f = E f e f )
by comparing it with the test results of the experimental part of this c= (9)
investigation. a1 fc′ b1b
Keywords: fiber-reinforced polymer; flexure; prestressing; post-tensioning;
in which c is the neutral axis depth of the section at nominal
strengthening; unbonded tendons.
flexural strength; Aps and dp are area and depth of the
unbonded prestressing steel (PS); Af is area and df is depth
INTRODUCTION AND BACKGROUND LITERATURE
of the FRP reinforcement; and As is area of bonded ordinary
The flexural capacity of bonded prestressed concrete
reinforcing steel (RS) at the section under consideration. For
members can be evaluated in accordance with the general
slabs, the areas Aps, As, and Af are per unit width b of the slab
ACI Building Code1 approach and the guidelines recom-
section. εce is precompression strain in concrete at the level
mended by ACI Committee 4402 by accounting for the effect
of the prestressed tendons; Ac and Ig are area and moment
of fiber-reinforced polymer (FRP) reinforcement as follows
of inertia of the gross section, respectively; e is eccentricity
of the tendons; Ec, Es, and Ef are modulus of elasticity of
( ) (
M n = Aps f ps d p − b1c / 2 + As fs ( d − b1c / 2 ) + ψ f A f E f e f d f − b1c / 2 ) (1) concrete, steel, and FRP reinforcement, respectively; εf and ff
are strain and stress in the FRP reinforcement, respectively;
where εfu is rupture strain of the FRP reinforcement; F is the mate-
rial stress-strain relationship of the prestressing reinforce-
fps = F(eps) (2) ment; εpe, fse, εps, and fps are effective strain, effective stress,
strain, and stress at ultimate, respectively, in the prestressing
steel (εpe = fse/Eps, where Eps is the modulus of elasticity of
fs = Eses ≤ fy (3) the prestressing steel); εs, fs, and fy are strain, stress, and
yield stress, respectively, for the ordinary RS; fc′ is concrete
cylindrical compressive strength; εcu is usable concrete
 dp − c  strain at compression failure (equal to 0.003 in accordance
e ps = e pe + e ce + e cu  (4)
 c  with ACI 318-111); α1 = 0.85 corresponding to a concrete
strain in the outermost compression fiber equal to εcu; β1 is
concrete strength factor defined in section 10.2.7.3 of the
ACI Building Code1; εbi is initial substrate strain at which
1  Aps fse Aps fse e 
2
the FRP was applied for strengthening (ACI Committee
e ce =  +  (5)
Ec  Ac Ig 
ACI Structural Journal, Vol. 111, No. 2, March-April 2014.
MS No. S-2012-168.R2, doi:10.14359.51686565, was received January 10, 2013,
and reviewed under Institute publication policies. Copyright © 2014, American
Concrete Institute. All rights reserved, including the making of copies unless
permission is obtained from the copyright proprietors. Pertinent discussion including
author’s closure, if any, will be published ten months from this journal’s date if the
discussion is received within four months of the paper’s print publication.

ACI Structural Journal/March-April 2014 407


4402), determined using elastic cracked section analysis,
considering all loads that will be on the member during the
FRP installation; εfd is strain in the FRP reinforcement at
which FRP debonding failure occurs; nf and tf are number of
FRP layers and thickness per one layer, respectively; and ψf
is the FRP strength-reduction factor recommended by ACI
Committee 440,2 which is taken equal to 0.85 for flexure.
To calculate the nominal moment capacity of
FRP-strengthened reinforced concrete (RC) or bonded PC
members, ACI Committee 4402 recommends using a trial-
and-error procedure for estimating the neutral axis depth
c until the requirements of strain compatibility and force
equilibrium across the depth of the critical section are satis-
fied (Eq. (9)). Also, two modes of flexural failure are recog-
nized by ACI Committee 440: 1) concrete crushing—that is,
when the strain in the outermost concrete compression fiber
reaches εcu before FRP failure; and 2) FRP failure before
concrete crushing. FRP failure could occur either by FRP
rupture, cover delamination, or FRP debonding. Accord- Fig. 1— FRP-strengthened continuous unbonded member
ingly, ACI Committee 4402 limits, conservatively, the strain with multi-collapse mechanisms.
in the FRP reinforcement at which FRP failure takes place to
the debonding strain εfd (Eq. (8)). When the strain in the FRP ACI Committee 4402 guidelines for calculating Mn for FRP
reaches its limiting strain εfd before the concrete compres- strengthened bonded PC or RC members. The approach is
sion strain reaches εcu, the concrete compression strain εcu validated for simply supported members by comparing with
in Eq. (2) through (9) should be replaced by its actual value the test results generated in the experimental part of this
expressed as a function of εfd as follows investigation.

 c  STRESS IN UNBONDED TENDONS AT ULTIMATE


( )
e c = e fd + e bi  
 df − c
(10) While the approach recommended by ACI Committee
4402 for estimating Mn for FRP strengthened RC or bonded
PC members is simple because it relies on commonly
The values of α1 and β1, which are now different from adopted and familiar principles, it is not as straightforward
their values when εc = εcu, can be estimated with reasonable for PC members with unbonded tendons. In PC members
accuracy using the following expressions recommended by with internal or external unbonded tendons, because the
ACI Committee 440,2 which were derived assuming a para- unbonded steel slips relative to the surrounding concrete,
bolic relationship for the stress-strain curve of concrete in the strain or stress fps in the prestressing steel that develops
compression at nominal flexural strength relies on the deformation of
the member as a whole. This makes the evaluation of the
4 e c′ − e c corresponding strain and stress member, rather than section,
b1 = (11) dependent.
6 e c′ − 2 e c
Several design approaches are available for predicting
the stress in unbonded tendons at ultimate.1,3,4 A compara-
tive evaluation of these approaches has been discussed in
3e c′ e c − (e c )2 detail elsewhere.5 Using plastic analysis and the concept
a1 = (12)
3b1 (e c′ )2 of collapse mechanism in continuous members, as well as
idealization of the curvature distribution along the span
where εc′ is the strain corresponding to fc′, which can either lengths at nominal flexural strength as shown in Fig. 1 (for
be taken equal to 0.002 or calculated more accurately as εc = εcu, and Af = 0.0) and further incorporating an empir-
εc′ = 1.7fc′/Ec. ically derived expression for the equivalent plastic hinge
length for unbonded members given as5 lp = (20.7/f + 10.5)c,
RESEARCH SIGNIFICANCE Harajli6 developed the following general and yet elegant
An experimental study was carried out and a design- expression for calculating the strain εps in unbonded tendons
oriented approach was developed for evaluating the nominal at ultimate
flexural capacity Mn of unbonded PC members when
strengthened using external FRP composites. The design  dp c
approach builds on a previous model generated for evalu- e ps = e pe + f ps  N p e cu − N p (e cu − e ce )  (13)
 La La 
ating the stress in unbonded tendons at ultimate in simply
supported or continuous members, and is consistent with the
where

408 ACI Structural Journal/March-April 2014


 20.7 
Np =  + 10.5 n p + + 10.5n p − (14)
 f 

in which c is the neutral axis depth at the section under consid-


eration; εce is calculated using Eq. (5) by neglecting the effect
of the secondary moment due to prestressing in continuous
members; La is the total length of tendons between anchor-
ages; Np (Eq. (14)) is a parameter that combines the effect
of member continuity and type of applied load; f = ∞ for a
single concentrated load, 3.0 for two-third point loads, and
6.0 for uniform load application, respectively; np+, np– are the
number of positive and negative plastic hinges, respectively,
that develop in the process of forming a collapse mecha-
nism; and φps is a stress-reduction factor taken equal to 0.7.
Note that because the live load in buildings and bridges is
seldom a concentrated load, and also because the dead load
is uniformly distributed, the multiplier (20.7/f + 10.5) of the
positive number of plastic hinge(s) np+ can always be set
equal to 14.0 to correspond to uniform load application.6
Equations (13) and (14) were developed with the perspec-
tive that the calculation of the stress in unbonded tendons
at ultimate in continuous members differs from one critical
section to another. That is, the stress at a critical section Fig. 2—Loading pattern and corresponding values of conti-
depends on the pattern of applied load and the conse- nuity parameter Np for continuous members.
quent collapse mechanism that would potentially develop
supported or continuous members at nominal flexural
for producing the maximum factored design moment at
strength
that section. For the hypothetical case of Fig. 1, in which
collapse mechanisms are assumed to form in all spans:
np1+ = 2.0, np2– = 2.0, and np3+ = 2.0. In actual design,  dp − c 
e ps = e pe + f ps N p e c  (15)
however, the numbers of plastic hinges np+ and np– are  La 

obtained from the collapse mechanisms that develop when
loading the minimum number of spans for producing Recognizing that the stress in the prestressing steel
maximum moment at the section under consideration. For seldom exceeds yield and limiting the corresponding stress
instance, in simply supported members, one span is loaded, to 0.95fpy6 allows the use of a linear relationship between the
and hence np+ = 1.0, np– = 0.0, and Np = 14.0. For continuous stress and strain in the prestressing steel, that is, fps = Epsεps,
members, collapse mechanisms (a), (b), (c), (d), and (e) in leading to
Fig. 2, and corresponding values of Np are recommended6
for computing the tendon stress fps and the nominal moment
f ps N p E ps e c  c
capacity at the maximum positive moment section in exte- f ps = fse +  1 −  ≤ 0.95 f py (16)
rior spans; negative moment section at the interior support La / d p  d p 

of a two-span member; maximum positive moment section
in interior spans; negative moment section at the first inte-
The force equilibrium across the depth of the critical
rior support; and negative moment section at the remaining
section, assuming rectangular section or rectangular section
interior supports of members with more than two spans,
behavior, is expressed as
respectively.
Apsfps + Asfs + AfEfef = a1fc′bb1c (17)
PROPOSED APPROACH FOR COMPUTING MN IN
FRP STRENGTHENED UNBONDED MEMBERS
It should be noted that for unbonded members without
Using Eq. (13) but neglecting the precompression strain
FRP reinforcement (that is, Af = 0.0 and α1 = 0.85), the value
εce due to its minor effect on the tendon stress in unbonded
of c from Eq. (17) can be integrated in Eq. (16) to produce
members, particularly when compared with bonded
a direct expression for computing fps in unbonded members
members, and considering that the flexural strength may
at ultimate.6
be controlled by FRP failure at which εc ≤ εcu (Fig. 1), the
Using Eq. (15), (16), and (17), the following step-by-step
following expression is recommended for computing the
procedure can be adopted for evaluating the nominal moment
strain εps in unbonded tendons of FRP-strengthened simply
capacity Mn of FRP strengthened unbonded post-tensioned
members.

ACI Structural Journal/March-April 2014 409


Case I—Flexural capacity controlled by concrete Step 7—Repeat Steps 5 and 6 by revising the neutral axis
crushing depth c until the requirement of force equilibrium in Eq. (17)
Step 1—Assume that the nominal flexural strength at a is satisfied within some degree of tolerance.
critical section is controlled by concrete crushing—that is, εf Step 8—Calculate the nominal moment capacity Mn at the
calculated using Eq. (7) is less than or equal to εfd (Eq. (8)). section under consideration using Eq. (1) in which εf = εfd.
This implies that εc = εcu, α1 = 0.85, and β1is as defined in Step 9—Check if φMn ≥ Mu, where Mu is the load-factored
section 10.2.7.3 of the ACI Building Code.1 Replacing fps applied moment at the section under consideration obtained
from Eq. (16) into Eq. (17) and assuming that the RS yields, by loading the spans (pattern loading) for producing
that is, fs = fy, leads to the following quadratic equation for maximum moment at the section under consideration.
calculating the neutral axis depth at nominal flexural strength The strength-reduction factor φ is taken in accordance
at the section under consideration with the ACI Building Code1 approach using the relationship
between the net tensile strain in the tension reinforcement
B + B2 + 4 AC and the neutral axis depth c at nominal flexural strength as
c= (18) follows
2A
φ = 0.90 for c/de ≤ 0.38 (20a)
where

f ps N p Aps E ps e cu φ = 0.65 for c/de ≤ 0.6 (20b)


A = 0.85b1 fc′b + (19a)
La

 c
f = 0.65 + 0.25  2.73 − 4.55  for
 f ps N p E ps e cu d p   de 
(20c)
B = Aps  fse +
 La  + As f y − A f E f (e cu + e bi ) (19b)
0.38 ≤ c / de ≤ 0.6

where de is the equivalent depth of the tension reinforcement


C = Af Ef ecudf (19c) (PS, RS, and FRP laminates) expressed as

in which Np is as defined previously (Fig. 2). For simply ( Aps f ps d p + As fs d + A f E f e f d f )


supported members, Np = 14.0. de = (21)
Aps f ps + As fs + A f E f e f
Step 2—Check if εf (Eq. (7)) is indeed ≤ εfd. Also check
if the strain εs in the RS is larger than the yield strain εy. If
εf ≤ εfd while εs is less than εy, repeat Step 1 for recalcu-
lating more accurately the neutral axis depth using Eq. (18) EXPERIMENTAL STUDY
in which the coefficients A, B, and C are revised by substi- Twenty-four unbonded post-tensioned specimens were
 d − c tested to evaluate their nominal flexural strength before and
tuting fs = Es  e s = e cu c  for fy in Eq. (17). after FRP strengthening. An additional six bonded post-
Step 3—Calculate fps from Eq. (16) corresponding to tensioned and six RC specimens were also tested for
εc = εcu, and fs from Eq. (3), and hence calculate the nominal comparison. All 36 specimens were simply supported over
moment capacity Mn from Eq. (1). a 3.0 m (9.84 ft) span. Dimensions and reinforcement layout
are given in Fig. 3. Eighteen of the specimens had a 150 mm
Case II—Flexural capacity controlled by FRP (5.9 in.) wide by 250 mm (9.8 in.) deep cross section and
failure a span-depth ratio (depth to center of tension steel) of 15,
Step 4—If εf calculated from Step 2 is greater than εfd representing beam members, and the remaining 18 had
(Eq. (8)), then FRP failure occurs before the strain εc reaches a 360 mm (14.2 in.) wide by 120 mm (4.7 in.) deep cross
εcu. In this case, the strain εf in the FRP reinforcement is equal section and a span-depth ratio of 35, simulating one-way
to εfd, and hence, a trial-and-error procedure for calculating slab members. The specimen designation, along with areas
c, as described in the next steps, becomes more appropriate. and depths of reinforcement and other pertinent details
Step 5— Using the value of εf = εfd, together with an initial and design properties, are summarized in Table 1. The test
assumed value of c, calculate the concrete strain εc at the top parameters included, in addition to the span-depth ratio of
concrete compression fiber from Eq. (10) and calculate the the member, area of internal prestressed reinforcement for
stress in the prestressing steel (Eq. (16)) and the stress in the the PC specimens or area of ordinary tension reinforcement
RS from Eq. (3) and (6) by replacing εc for εcu. for the RC specimens; area of external FRP reinforcement;
Step 6—Check equilibrium of forces using Eq. (17) in and tendon profile. Two tendon profiles were selected for
which εf = εfd, and α1 and β1 are as calculated from Eq. (11) each set of specimens: one horizontal and one parabolic.
and (12). In the specimens designation provided in Table 1, the first
letter U stands for unbounded, B for bonded, and R for rein-
forced. The second letter B stands for beam, while S stands

410 ACI Structural Journal/March-April 2014


Fig. 3—Typical dimensions and reinforcement details of the beam and slab specimens. (Note: dimensions in mm; 1 mm = 0.039 in.)
for slab. The numbers 1 and 2 following the second letter where fpy = 1670 MPa (242 ksi), Eps = 195,130 MPa
designate two different levels of PS or RS areas or ratios. (28,400 ksi), N = 14.84, K = 1.0, and Q = 0.0357 for the
Letters H and P designate horizontal and parabolic tendon 7.9 mm (5/16 in.) strands; and fpy = 1690 MPa (245 ksi), Eps
profile, respectively, and F1 and F2 denote two different = 194,440 MPa (28,200 ksi), N = 12.1, K = 1.011, and Q =
levels (areas or layers) of external FRP reinforcement. The 0.0301 for the 9.5 mm (3/8 in.) strands.
parabolic tendon profile in all beam and slab specimens had All ordinary reinforcing bars (except the 6 mm [0.25 in.]
zero eccentricity at the support. bars) were deformed Grade 60 steel with actual yield strength
The prestressing reinforcement consisted of seven-wire as given in Table 1. The FRP composite used for strength-
strands having 7.9 and 9.5 mm (5/16 and 3/8 in.) diameter, ening consisted of carbon fiber-reinforced polymer (CFRP)
with an area of 37.5 and 52.0 mm2 (0.058 and 0.08 in.2), flexible sheets having unidirectional carbon fabric with glass
and actual ultimate strength fpu of 1958 and 1978 MPa (284 cross-fiber for added strength and fabric stability during
and 287 ksi), respectively. Generated from coupon tests, the installation. The design thickness, the modulus of elas-
actual stress-strain (fps – εps) behavior of the two sizes of the ticity, and the ultimate tensile strain of the fibers are 1 mm
prestressing steel were best reproduced using the following (0.039 in.), 95,800 MPa (13,895 ksi), and 1%, respectively.
relationship7 Before casting the specimens, the steel cages were
instrumented with electric strain gauges and then placed in
  plywood formwork ready for concrete casting. The ducts
  for the post-tensioned steel consisted of galvanized flexible
  tubes 20 mm (0.79 in.) in diameter. One strand was placed
1−Q
f ps = E ps e ps  Q + 
(22) in each duct. For the bonded post-tensioned slab and beam
   E e N 
1/ N
specimens, cement-based grout was injected inside the ducts
 1 +  ps ps   
   Kf py    after the tendons were stressed to provide bond between the
   strands and concrete. The grout mixture was prepared using

Type I portland cement, and proportioned in accordance

ACI Structural Journal/March-April 2014 411


Table 1—Summary of test parameters
Prestressing steel Reinforcing steel FRP
Af,
Concrete Specimen Tendon fse, As, As′, fy, mm2 fc′,
system label profile Aps dp, mm MPa bottom d,mm top Av MPa (nf) MPa
UB1-H Horizontal 1 (5/16 in.) 200 813 2 (8 mm) 220 — φ8 at 150 560 — 42
UB1-H-F1 Horizontal 1 (5/16 in.) 200 962 2 (8 mm) 220 — φ8 at 150 612 150 (1) 36
UB1-H-F2 Horizontal 1 (5/16 in.) 200 963 2 (8 mm) 220 — φ8 at 150 612 300 (2) 36
UB1-P Parabolic 1 (5/16 in.) 200 815 2 (8 mm) 220 — φ8 at 150 560 — 42
UB1-P-F1 Parabolic 1 (5/16 in.) 200 971 2 (8 mm) 220 — φ8 at 150 612 150 (1) 36
UB1-P-F2 Parabolic 1 (5/16 in.) 200 781 2 (8 mm) 220 — φ8 at 150 612 300 (2) 37
Beam

UB2-H Horizontal 2 (3/8 in.) 200 778 2 (8 mm) 220 2 (8 mm) φ8 at 150 560 — 42
UB2-H-F1 Horizontal 2 (3/8 in.) 200 924 2 (8 mm) 220 2 (8 mm) φ8 at 150 612 150 (1) 36
UB2-H-F2 Horizontal 2 (3/8 in.) 200 896 2 (8 mm) 220 2 (8 mm) φ8 at 150 612 300 (2) 37
UB2-P Parabolic 2 (3/8 in.) 200 836 2 (8 mm) 220 2 (8 mm) φ8 at 150 560 — 42
Unbonded post-tensioned

UB2-P-F1 Parabolic 2 (3/8 in.) 200 936 2 (8 mm) 220 2 (8 mm) φ8 at 150 612 150 (1) 36
UB2-P-F2 Parabolic 2 (3/8 in.) 200 923 2 (8 mm) 220 2 (8 mm) φ8 at 150 612 300 (2) 37
US1-H Horizontal 2 (5/16 in.) 85 927 2 (8 mm) 92.5 — — 560 — 42
US1-H-F1 Horizontal 2 (5/16 in.) 85 917 2 (8 mm) 92.5 — — 612 150 (1) 36
US1-H-F2 Horizontal 2 (5/16 in.) 85 964 2 (8 mm) 92.5 — — 612 300 (1) 36
US1-P Parabolic 2 (5/16 in.) 85 886 2 (8 mm) 98.5 — — 560 — 42
US1-P-F1 Parabolic 2 (5/16 in.) 85 949 2 (8 mm) 98.5 — — 612 150 (1) 36
US1-P-F2 Parabolic 2 (5/16 in.) 85 971 2 (8 mm) 98.5 — — 612 300 (1) 37
Slab

US2-H Horizontal 3 (3/8 in.) 85 804 2 (8 mm) 92.5 — — 560 — 42


US2-H-F1 Horizontal 3 (3/8 in.) 85 912 2 (8 mm) 92.5 — — 612 150 (1) 36
US2-H-F2 Horizontal 3 (3/8 in.) 85 858 2 (8 mm) 92.5 — — 612 300 (1) 37
US2-P Parabolic 3 (3/8 in.) 85 831 2 (8 mm) 98.5 — — 560 — 42
US2-P-F1 Parabolic 3 (3/8 in.) 85 921 2 (8 mm) 98.5 — — 612 150 (1) 36
US2-P-F2 Parabolic 3 (3/8 in.) 85 916 2 (8 mm) 98.5 — — 612 300 (1) 37
BB2-P Parabolic 2 (3/8 in.) 200 884 2 (6 mm) 220 2 (6 mm) φ8 at 150 0 — 37
Bonded post-tensioned

Beam

BB2-P-F1 Parabolic 2 (3/8 in.) 200 894 2 (6 mm) 220 2 (6 mm) φ8 at 150 0 150 (1) 37
BB2-P-F2 Parabolic 2 (3/8 in.) 200 885 2 (6 mm) 220 2 (6 mm) φ8 at 150 0 300 (2) 37
BS2-P Parabolic 3 (3/8 in.) 85 970 2 (8 mm) 98.5 — — 0 — 37
Slab

BS2-P-F1 Parabolic 3 (3/8 in.) 85 915 2 (8 mm) 98.5 — — 0 150 (1) 37


BS2-P-F2 Parabolic 3 (3/8 in.) 85 892 2 (8 mm) 98.5 — — 0 300 (1) 37
RB2 — — — — 2 (16 mm) 220 — φ8 at 100 530 — 37
Beam
Reinforced concrete

RB2-F1 — — — — 2 (16 mm) 220 — φ8 at 100 530 150 (1) 37


RB2-F2 — — — — 2 (16 mm) 220 — φ8 at 100 674 300 (2) 37
RS2 — — — — 4 (12 mm) 100 — — 555 — 37
Slab

RS2-F1 — — — — 4 (12 mm) 100 — — 555 150 (1) 37


RS2-F2 — — — — 4 (12 mm) 100 — — 624 300 (1) 37

Note: 1 in. = 25.4 mm; 1 MPa = 0.145 ksi; 1 mm2 = 0.0016 in.2.

with the ACI Building Code1 with a water-cement ratio of the manufacturer’s recommendations and in compliance
0.40. with the ACI Committee 4402 recommendation for securing
The CFRP sheets were attached to the bottom tension proper development length. No particular measures were
face of the beam and slab specimens in accordance with taken to improve bond strength between the FRP and the

412 ACI Structural Journal/March-April 2014


substrate. One or two layers of 150 mm (5.9 in.) wide FRP
sheets were applied for the beam specimens, while only one
layer of 150 or 300 mm (5.9 or 11.8 in.) wide sheets was
applied for the slab specimens (Table 1).
All specimens were tested in four-point bending using two
symmetrical concentrated loads spaced a distance equal to
1/6 the span length, or 500 mm (19.7 in.) apart (Fig. 3). It
should be noted that because the increase in tendon strain/
stress (above effective prestrain/prestress) of unbonded
members due to increase in applied load depends on the
overall deformation of the member or elongation of the
tendon between the anchorages, the ultimate tendon stress
and flexural capacity of unbonded members are influenced
by the geometry of applied load.8 Two-point loads spaced at
1/6 the span length were found analytically8,9 to be equiva-
lent to uniform load application for predicting the ultimate
tendon stress and moment capacity of unbonded post-ten-
sioned members.
To replicate actual conditions of concrete flexural
members that require strengthening or repair, all control
and strengthened specimens were first subjected to cyclic
loading consisting of six cycles before the FRP applica-
tion, and another six cycles after FRP application, ranging
between a minimum load Pmin and a maximum load Pmax,
simulating dead load and dead plus live load, respectively.
The loads Pmin and Pmax were set at 30 and 70%, respectively,
of the calculated nominal load capacity of the specimens.
For the control or unstrengthened specimens, the cyclic
loading stage was followed immediately by a stage of mono-
tonically increasing load until complete flexural failure of
the specimens. The specimens that were strengthened
using CFRP were first subjected to the same cyclic loading
protocol as the control specimens, and then unloaded to
prepare them for CFRP application. Following at least
7 days of CFRP application, the strengthened specimens Fig. 4—Typical photos of specimens at conclusion of test.
were subjected to a loading protocol consisting of cyclic steel, strain in the CFRP reinforcement, and mode of failure
loading and monotonically increasing load to failure, similar are all summarized in Table 2. Representative variations of
to the control specimens. deflection, FRP strain, and stress increase in the prestressing
Test measurements included strains and stresses in the steel above effective prestress fse with applied load are shown
prestressing strands, CFRP laminates, and reinforcing bars in Fig. (6) through (9), respectively.
of the RC specimens within the constant moment region The crack patterns for the unbonded beam specimens
close to midspan; applied load; and deflection. Crack (Fig. 5) and the slab specimens (not shown for brevity) were
patterns were also monitored throughout the test for each similar to those developed in the companion bonded speci-
specimen. The strains were measured using electric strain mens. The spreading of cracks outside the constant moment
gauges, while deflection was measured using a linear voltage region increased with the use of FRP reinforcement and
differential transformer (LVDT). The test results were auto- as the area of FRP reinforcement increased, and was most
matically collected and recorded using a data acquisition and significant for the RC specimens (Fig. 5).
control system. It is clear from the test data summarized in Table 2 that
the use of FRP reinforcement significantly increased the
DISCUSSION OF RELEVANT TEST RESULTS moment capacity of unbonded post-tensioned members. The
Typical photos of the specimens at the conclusion of the corresponding increase grew higher as the area of FRP rein-
test are shown in Fig. 4. Representative modes of failure forcement increased, and varied between 24 and 105% for
and cracking patterns at nominal strength for the unbonded the beam specimens and between 21 and 126% for the slab
PC beam specimens in comparison with the bonded PC and specimens of the current investigation. As would be expected,
RC beam specimens are provided in Fig. 5. Flexural failure however, the increases in load capacities were accompanied
for the various beam and slab specimens occurred either with reductions in ductility or ultimate deformation capaci-
by concrete crushing or by FRP debonding or fracturing. ties, which were most notable for the beam specimens. The
Relevant experimental results at nominal flexural strength reductions in deformation capacity, which are attributed in
including load capacity, deflection, stress in the prestressing

ACI Structural Journal/March-April 2014 413


were significantly larger than the stresses developed in the
unbonded specimens.
Based on the experimental results and observations of
the current investigation, it was obvious that the cracking
and crack patterns, increase in load capacities, reduction in
deformation capacities, and modes of flexural failure of the
unbonded PC specimens as a result of FRP strengthening
were quite similar to those of the bonded PC and RC speci-
mens. More test data and a detailed discussion of test results
are reported elsewhere.10

COMPARISON OF ANALYTICAL PREDICTIONS


WITH TEST DATA
Table 2 shows comparisons of the predictions of the
proposed design-oriented approach developed in this study
for FRP-strengthened unbonded PC members together with
the predictions of the ACI Committee 4402 approach for
bonded PC and RC members against the test results gener-
ated in the experimental part of this investigation. These
include stress fps in the prestressing steel, strain εf in the FRP
reinforcement, mode of flexural failure, and nominal moment
capacity Mn. In calculating the stress in the prestressing steel
and the nominal moment capacity, the reduction factors φps
and ψf were both set equal to 1.0. Also, because the load at
which the FRP was applied was very small (equal to the self-
weight of the specimen), the initial substrate strain εbi was
taken equal to zero.
Figure 10 shows predicted stress increase Δfps (Δfps = fps – fse)
and total stress fps in the prestressing steel at ultimate versus
test results for the unbonded prestressed beams and slab
specimens of the current investigation. The data are super-
imposed on predicted versus test results of a collection of
unbonded members (without FRP reinforcement) compiled
and reported previously.5,6 The compiled data correspond
to internally or externally post-tensioned simply supported
members, and continuous members having two or three
Fig. 5—Typical comparison of crack pattern for unbonded spans, and loaded with two-point or single-point loads,
PC, bonded PC, and RC beam specimens. respectively. Figure 11 shows predicted versus measured
nominal moment capacities for the 36 beam and slab speci-
part to the FRP debonding mode of flexural failure, were
mens tested in this investigation.
almost identical for the slab or beam specimens reinforced
Given the inevitable scatter associated with the prediction
with Level F1 and F2 FRP reinforcement (Table 2).
of the stress in unbonded tendons at ultimate, it can be seen
The FRP strain/stress developed at flexural failure (Fig. 8
from Fig. 10 that the proposed analytical approach predicted
and Table 2) generally decreased as the area of FRP rein-
the test data with reasonable accuracy. More conserva-
forcement increased within the specimens of the same test
tive predictions can be obtained by using the stress reduc-
series, or as the area of prestressing reinforcement increased.
tion parameter φps = 0.7 proposed for design purposes. As
The overall average FRP strains at ultimate for the combined
expected, the stress predictions for the bonded members
beam and slab specimens strengthened using the two
were more accurate than the unbonded members. The
different areas or levels (F1 and F2) of FRP reinforcement
average ratio of the test-to-predicted tendon stress were 1.10
were 6260 to 5400 με for the unbonded PC, 6875 to 4970 με
(standard deviation [SD] = 0.12) for the unbonded speci-
for the bonded PC, and 7608 to 5419 με for the RC speci-
mens, and 1.02 (SD = 0.04) for the bonded specimens. It
mens, respectively.
should be noted that due to the presence of 2φ8 mm ordi-
Being unbonded, the stress in the prestressing steel at
nary steel bars, which are required as minimum bonded rein-
nominal flexural strength was below yield for all specimens.
forcement in accordance with the ACI Building Code,1 the
The corresponding strain (Table 2) decreased as the area of
control unbonded PC specimens developed well-distributed
FRP reinforcement increased. On the other hand, the strains
cracks along their length as opposed to the development of
and stresses in the prestressing steel for all companion
few cracks or concentration of deformation at a single crack
bonded PC beam and slab specimens exceeded yield and
that normally occurs in members with an unbonded tendon
system.5 In other words, the equivalent plastic hinge length

414 ACI Structural Journal/March-April 2014


Table 2—Summary of test results and analytic predictions
Pu Δu
(kN) (mm) fps (MPa) FRP Strain εf Mode of failure Nominal moment Mn (kN-m)
Exp./ Exp./ Exp./
Specimen Exp. Exp. Exp. Analysis Analysis Exp. Analysis Analysis Experiment Analysis Exp. Analysis Analysis
UB1-H 42.3 64 1567 1253 1.25 — — — Concrete crushing Concrete crushing 26.4 20.8 1.27
UB1-H-F1 66.9 31 1303 1250 1.04 7122 7948* 0.90 FRP debonding FRP debonding 41.8 46.5 0.90
UB1-H-F2 86.9 33 1223 1290 0.95 5378 5620* 0.96 FRP debonding FRP debonding 54.3 55.6 0.98
UB1-P 46.8 81 1669 1255 1.33 — — — Concrete crushing Concrete crushing 29.3 20.8 1.41
UB1-P-F1 66.3 35 1413 1259 1.12 4556 7948* 0.57 FRP debonding FRP debonding 41.4 46.5 0.89
*
UB1-P-F2 89.0 36 1102 1098 1.00 5604 5698 0.98 FRP debonding FRP debonding 55.6 55.2 1.01
UB2-H 63.6 43 1246 1205 1.03 — — — Concrete crushing Concrete crushing 39.8 35.3 1.13
Partial rupture +
UB2-H-F1 80.8 26 1163 1230 0.95 6934 7949* 0.87 FRP debonding 50.5 60.4 0.84
partial debonding
UB2-H-F2 104.8 31 1122 1239 0.91 5329 5698* 0.94 FRP debonding FRP debonding 65.5 68.9 0.95
UB2-P 75.2 88 1598 1260 1.27 — — — Concrete crushing Concrete crushing 47.0 36.3 1.29
FRP debonding +
UB2-P-F1 93.6 37 1339 1244 1.08 5393 7948* 0.68 FRP debonding 58.5 60.6 0.97
concrete crushing
UB2-P-F2 101.2 30 1223 1270 0.96 4285 5698* 0.75 FRP debonding FRP debonding 63.3 69.3 0.91
US1-H 22.7 62 1211 1106 1.09 — — — Concrete crushing Concrete crushing 14.2 11.5 1.23
*
US1-H-F1 34.3 63 1195 1033 1.16 8309 7948 1.05 FRP rupture FRP debonding 21.4 23.3 0.92
Partial rupture +
US1-H-F2 43.0 66 1152 1112 1.04 6074 6490 0.94 Concrete crushing 26.9 30.1 0.89
partial debonding
US1-P 21.3 100 1413 1066 1.33 — — — Concrete crushing Concrete crushing 13.3 11.6 1.15
US1-P-F1 34.5 68 1177 1065 1.11 5770 7948* 0.73 FRP debonding FRP debonding 21.6 23.8 0.91
FRP rupture +
US1-P-F2 48.1 75 1165 1117 1.04 6280 6580 0.95 Concrete crushing 30.1 30.8 0.97
concrete crushing
US2-H 35.1 87 1227 966 1.27 — — — Concrete crushing Concrete crushing 21.9 16.3 1.34
Concrete crushing
US2-H-F1 42.6 62 — 1039 — 6277 6758 0.93 Concrete crushing 26.6 26.4 1.01
+ partial debonding
US2-H-F2 57.2 75 1065 972 1.10 5517 5551 0.99 Concrete crushing Concrete crushing 35.8 31.6 1.13
US2-P 36.9 66 1105 992 1.11 — — — Concrete crushing Concrete crushing 23.1 17.0 1.36
US2-P-F1 47.6 68 1146 1047 1.09 7554 6731 1.12 Concrete crushing Concrete crushing 29.8 26.9 1.11
US2-P-F2 59.8 67 1136 1029 1.10 5570 5429 1.03 Concrete crushing Concrete crushing 37.4 32.1 1.16
BB2-P 63.4 44 1738 1737 1.00 — — — Concrete crushing Concrete crushing 39.6 34.2 1.16
BB2-P-F1 87.0 35 1710 1682 1.02 5906 7027 0.84 FRP rupture Concrete crushing 54.4 53.6 1.01
BB2-P-F2 107.8 30 1680 1591 1.06 4781 5523* 0.87 FRP debonding FRP debonding 67.4 63.1 1.07
BS2-P 33.9 71 1662 1687 0.99 — — — Concrete crushing Concrete crushing 21.2 21.4 0.99
FRP debond. and
BS2-P-F1 44.3 67 1702 1556 1.09 7834 6000 1.31 Concrete crushing 27.7 28.1 0.98
rupture
BS2-P-F2 59.7 70 1429 1435 1.00 5701 5045 1.13 Concrete crushing Concrete crushing 37.3 32.2 1.16
RB2 72.5 46 — — — — — — Concrete crushing Concrete crushing 45.3 42.5 1.07
RB2-F1 98.6 35 — — — 7608 6740 1.13 FRP rupture Concrete crushing 61.6 62.0 0.99
RB2-F2 110.8 33 — — — 5003 4177 1.20 Concrete crushing Concrete crushing 69.3 73.4 0.94
RS2 37.0 77 — — — — — — Concrete crushing Concrete crushing 23.1 22.7 1.02
RS2-F1 48.5 60 — — — — 6863 — FRP rupture Concrete crushing 30.3 32.2 0.94
RS2-F2 66.7 77 — — — 5834 4660 1.25 FRP rupture Concrete crushing 41.7 36.6 1.14
*
Equal to debonding strain calculated using Eq. (8).
Notes: Exp. is Experiment; 1 kN = 0.224 kip; 1 mm = 0.039 in.; 1 MPa = 0.145 ksi; 1 kN-m = 8.83 k-in.)

ACI Structural Journal/March-April 2014 415


Fig. 7—Representative load-deflection response of slab
specimens. (Note: 1 mm = 0.039 in.; 1 kN = 0.224 kip.)

Fig. 6—Comparison between load-deflection response of


unbonded and bonded PC beam specimens. (Note: 1 mm =
0.039 in.; 1 kN = 0.224 kip.)
developed in these specimens was, admittedly, larger than
that predicted from lp = (20.7/f + 10.5)c, based on which
Eq. (13) and (16) were developed. Consequently, the ulti-
mate strains or stresses in the prestressing steel for the Fig. 8—Representative variation of FRP strain with applied
control specimens in this investigation (Table 2 and Fig. 10) load. (Note: 1 kN = 0.224 kip.)
were conservatively larger than those predicted by Eq. (13)
or (16).
Also, as can be seen from Table 2, the analytical approach
predicted the experimentally measured strain in the FRP
reinforcement and associated mode of flexural failure
reasonably accurately for most unbonded specimens, as well
as for the bonded PC and RC specimens. While the results
are, to a great extent, in support of the ACI Committee
4402 guidelines for predicting modes of flexural failure in
FRP-strengthened members, the experimentally measured
FRP strains for the unbonded specimens were consistently
slightly lower than those predicted using Eq. (8). The
average ratio of test-to-predicted FRP strains is calculated at
0.9 (SD = 0.15) for the unbonded specimens, 1.1 (SD = 0.18)
for the combined bonded PC and RC specimens, and 0.96
(SD = 0.18) for all unbonded, bonded, and RC specimens.
Finally, from Fig. 11, Table 2, and the statistical data
provided, except for the control unbonded PC specimens
that developed larger than predicted moment capacities due
to the development of larger-than-predicted steel stresses, Fig. 9—Typical variation of stress increase in unbonded
prestressing steel with applied load. (Note: 1 kN = 0.224 kip.)

416 ACI Structural Journal/March-April 2014


Fig. 11—Prediction of nominal moment capacity of combined
specimens of current investigation. (Note: 1 kN-m = 8.83 k-in.)
specimens, particularly when the mode of failure is by
FRP debonding or rupture, were slightly lower than those
predicted by Eq. (8), these specimens developed slightly
lower moment capacities than those predicted using the
proposed approach (Table 2). Consequently, although addi-
tional conservatism in predicting Mn for design purposes
can be gained by setting ψf = 0.85 as recommended by
ACI Committee 4402 (Eq. (1)), the FRP debonding strain
in Eq. (8) may require slight modification in its application
for unbonded members. Until further experimental evidence
is available to support such modification, however, Eq. (8)
can still be used for unbonded members without significant
Fig. 10—Prediction of experimental data for unbonded
loss of accuracy.
members6 including FRP strengthened unbonded specimens of
current investigation for φps = 1.0. (Note: 1 MPa = 0.145 ksi.)
CONCLUSIONS
as illustrated previously, the predicted nominal moment A design-oriented approach for calculating the nominal
capacities Mn of the FRP-strengthened specimens (assuming moment capacity Mn of unbonded PC members when
ψf = 1.0) were generally in good agreement with the test strengthened using external FRP composites is presented.
results. Despite little discrepancy, the level of accuracy The approach is applicable for simply supported and contin-
in predicting Mn for the unbonded specimens using the uous members, and is consistent with the guidelines of
proposed design approach is consistent with the level of ACI Committee 4402 for evaluating Mn of bonded PC or
accuracy in predicting Mn for bonded PC and RC specimens RC members. The accuracy of the proposed approach for
using the ACI Committee 4402 approach. The average ratio simply supported members was verified by comparing with
of test-to-predicted results is calculated as 0.97 (SD = 0.09) the results of a comprehensive test program designed and
for the FRP-strengthened unbonded specimens, and 1.05 carried out specifically for the purpose of this study.
(SD = 0.09) for the combined FRP-strengthened bonded Except for a slight discrepancy in predicting the FRP
PC and RC specimens. For the combined control and debonding strain for the unbonded specimens that encoun-
FRP-strengthened specimens, the coefficient of correlation tered FRP failure before concrete crushing, the proposed
R between the experimentally measured and the calculated approach for unbonded PC members, and the approach
nominal moment capacities is equal to 0.97 for the unbonded recommended by ACI Committee 440 for bonded PC and RC
specimens, 0.98 for the bonded PC and RC specimens, and members, predicted well the test results at ultimate, including
0.97 for the overall unbonded PC, bonded PC, and RC the strain/stress in the PS, strain in the FRP reinforcement,
specimens. Their corresponding average ratios of test-to- mode of flexural failure, and nominal flexural strength.
predicted results are equal to 1.07 (SD = 0.17), 1.04 (SD = The experimental results and the proposed approach for
0.08), and 1.06 (SD = 0.15). calculating Mn of FRP-strengthened unbonded PC members
It should be noted that because the experimentally clearly show that the use of external FRP reinforcement is
measured FRP strains of the FRP-strengthened unbonded as effective in improving the nominal flexural strength of

ACI Structural Journal/March-April 2014 417


unbonded PC members, as when used for strengthening 440.2R-08),” American Concrete Institute, Farmington Hills, MI, 2008,
76 pp.
bonded PC or RC members. Consequently, in designing 3. AASHTO, “LRFD Bridge Design Specifications,” American Asso-
the FRP system for flexural strengthening of unbonded PC ciation of State Highway and Transportation Officials, Washington, DC,
members, no special guidelines are needed beyond those 2004, 1324 pp.
4. Naaman, A. E.; Burns, N.; French, C.; Gamble, W. L.; and Mattock, A.
recommended in the ACI Committee 4402 report and the H., “Stresses in Unbonded Prestressing Tendons at Ultimate: Recommen-
design-oriented approach proposed for unbonded members dation,” ACI Structural Journal, V. 99, No. 4, July-Aug. 2002, pp. 518-529.
in this investigation. 5. Harajli, M. H., “On the Stress in Unbonded Tendons at Ultimate: Crit-
ical Assessment and Proposed Changes,” ACI Structural Journal, V. 103,
No. 6, Nov.-Dec. 2006, pp. 803-812.
AUTHOR BIOS 6. Harajli, M. H., “Tendon Stress at Ultimate in Continuous Unbonded
Fatima El Meski is a Lecturer at the Lebanese American University, Post-Tensioned Members: Proposed Modification of ACI Eq. (18-4)
Beirut, Lebanon. She was formerly a PhD Student in the Department of and (18-5),” ACI Structural Journal, V. 109, No. 2, Mar.-Apr. 2012, pp.
Civil and Environmental Engineering at the American University of Beirut, 183-192.
Beirut, Lebanon. 7. Menegotto, M., and Pinto, P. E., “Method of Analysis for Cyclically
Loaded Reinforced Concrete Plane Frames.” IABSE Preliminary Report for
ACI member Mohamed H. Harajli is a Professor of civil engineering at Symposium on Resistance and Ultimate Deformability of Structures Acted
the American University of Beirut. His research interests include design on Well-Defined Repeated Loads, Lisbon, Portugal, 1973, pp. 15-22.
and behavior of reinforced, prestressed, and fiber-reinforced concrete 8. Harajli, M. H., and Hijazi, S., “Evaluation of the Ultimate Steel Stress
members and strengthening and repair of concrete structures. in Unbonded Partially Prestressed Beams,” PCI Journal, V. 36, No. 2,
Jan.-Feb. 1991, pp. 62-82.
REFERENCES 9. Moon, J. H., and Burns, N. H., “Flexural Behavior of Members with
1. ACI Committee 318, “Building Code Requirements for Reinforced Unbonded Tendons. II: Application,” Journal of Structural Engineering,
Concrete and Commentary (ACI 318-11),” American Concrete Institute, ASCE, V. 123, No. 8, Aug. 1997, pp. 1095-1101.
Farmington Hills, MI, 2011, 503 pp. 10. El Meski, F., “Behavior of Unbonded Post Tensioned Members
2. ACI Committee 440, “Guide for the Design and Construction of Exter- Strengthened Using External FRP Composites: Experimental Evaluation
nally Bonded FRP Systems for Strengthening Concrete Structures (ACI and Analytic Modeling,” PhD dissertation, Department of Civil and Envi-
ronmental Engineering, American University of Beirut, Beirut, Lebanon,
2012, 283 pp.

418 ACI Structural Journal/March-April 2014


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 111-S37

Size Effect on Strand Bond and Concrete Strains at


Prestress Transfer
by José R. Martí-Vargas, Libardo A. Caro, and Pedro Serna-Ros
The size-effect phenomenon has been analyzed with pretensioned
prestressed concrete prismatic specimens at prestress transfer. An
experimental program that includes variables such as concrete
mixture design, specimen cross section size, and concrete age, has
been conducted. Several series of specimens with different embed-
ment lengths have been made and tested using a testing technique
based on the bond behavior analysis by measuring prestressing
strand force. In addition, some specimens have been instrumented
to determine longitudinal concrete strain profiles. The tests have
provided data on concrete strains, transfer length, effective
prestressing force, bond stress, and concrete modulus of elasticity.
Relationships between measured prestressing strand forces and
effective prestressing forces obtained from concrete compressive
strains have been found. A coefficient to account for the spec-
Fig. 1—Idealized prestressing strand force diagram.
imen crosssection size-effect on the concrete modulus of elasticity
has been proposed. Comparisons between test results and theo- assurance procedures for bonded applications should be
retical predictions from pre-existing equations in the codes have used.2 However, neither codes2,8,9 nor standards10,11 specify
been made.
minimum requirements for the bond performance of
Keywords: bond; modulus of elasticity; prestress; size-effect; strain; prestressing strands, and there is no consensus on the main
strand; transfer length. parameters to be considered in equations to compute transfer
length12,13 or on a standard test method for bond quality.3
INTRODUCTION By way of example, the ACI 318-112 code provisions
Force in a prestressing strand is transferred to concrete for transfer length are not a function of concrete strength,
by bond in the case of pretensioned prestressed concrete while the fib Model Code 20109 includes concrete proper-
members.1 At prestress transfer, the prestressing strand ties; moreover, several test procedures to determine bond
tends to shorten, the concrete around the prestressing strand characteristics14,15 or to measure transfer length16,17 have
shortens as the prestressing force is applied to it, and the been proposed as alternatives to the two most widely used
prestressing strand that is bonded to the concrete shortens methods18: measuring the strand end slip or determining
with it. Consequently, prestress loss due to the elastic short- the longitudinal concrete surface-strain profile. At prestress
ening of concrete occurs in the central zone of the member. transfer, variation in strand stress along the transfer length
At the member ends, the prestressing strand force varies from involves slips between the strand and the surrounding
zero to the effective prestressing force along the distance concrete. These slips can be used to estimate transfer
defined as transfer length2 (Fig. 1). length,19,20 but the results obtained from transfer length esti-
Both bond strength and transfer length depend on several mations vary vastly.21 After prestress transfer, the longitu-
factors, such as concrete strength at prestress transfer, initial dinal concrete surface strain profile also follows a similar
strand stress, concrete cover, prestress transfer method, law to that of the prestressing strand force shown in Fig. 1.
strand geometry, and strand surface condition,3 and bond Transfer length can be determined directly from the concrete
strength improves when a confining stress is applied.4,5 strain profile18,22,23 by either the Slope-Intercept Method24
Average bond stress along the transfer length has been (by intersecting an adjusted line to the initial branch with
characterized as being proportional to the square root of an adjusted line to the central branch) or the 95% Average
the concrete compressive strength given the influence Maximum Strain (AMS) Method7 (by intersecting the initial
of the elastic modulus of the concrete that surrounds the branch with the horizontal line corresponding to 95% of the
prestressing strand.6 average strains in the central branch).
A short transfer length results in higher stresses and Based on test results obtained from a recently devel-
risk of cracking near member ends, and a long transfer oped experimental methodology, the ECADA (Ensayo para
length shortens the available member length to resist the
ACI Structural Journal, V. 111, No. 2, March-April 2014.
bending moment and shear.6 As the transfer length is an MS No. S-2012-174.R2, doi:10.14359.51686530, was received February 19. 2013,
and reviewed under Institute publication policies. Copyright © 2014, American
important parameter in the design exercise,6,7 bond perfor- Concrete Institute. All rights reserved, including the making of copies unless
mance is assumed essential for an adequate response of permission is obtained from the copyright proprietors. Pertinent discussion including
author’s closure, if any, will be published ten months from this journal’s date if the
pretensioned prestressed concrete applications, and quality discussion is received within four months of the paper’s print publication.

ACI Structural Journal/March-April 2014 419


Table 1—Concrete mixture designs
Designation A B C
3 3
Cement, lb/yd (kg/m ) 843 (500) 624 (370) 674 (400)
Water-cement ratio (w/c) 0.3 0.45 0.5
Aggregate 7/12, lb/yd3 (kg/m3) 1674 (993) 1637 (971) 1645 (976)
Sand 0/4, lb/yd3 (kg/m3) 1468 (871) 1448 (859) 1443 (856)
High-range water-reducing admixture (%) 1.7 1.4 0.1
additive Type Polycarboxylate ether polymers Modified polycarboxylic ethers Modified polycarboxylic ethers
12 hours 5.8 (40) — —

Concrete compressive 24 hours 7.5 (52) 4.6 (32) 3.5 (24)


strength, ksi (MPa) 48 hours 8.4 (58) 5.2 (36) 4.2 (29)
28 days 12.3 (85) 7.3 (50) 6.3 (43)

Note: 1 mm = 0.04 in.

Caracterizar la Adherencia mediante Destesado y Arran- strengths at the time of testing (fci′), ranging from 24 to 55
camiento; in English: Test to Characterize the Bond by MPa (3.5 to 8.4 ksi), were tested. For all the concretes, the
Release and Pull-Out) test method,17 the research herein components were: cement CEM I 52.5 R,31 crushed lime-
analyzes the specimen cross section size-effect on strand stone aggregate 7 to 12 mm (0.275 to 0.472 in.), washed
bond and concrete strains at prestress transfer. Three rolled limestone sand 0 to 4 mm (0 to 0.157 in.), and a high-
different concrete mixture designs applicable to the precast range water-reducing admixture additive. The mixtures of
prestressed concrete members industry, in combination with the tested concretes are shown in Table 1.
three different specimen cross sections, have been tested. The prestressing strand was a low-relaxation seven-wire
steel strand specificied as UNE 36094:97 Y 1860 S7 13.010
RESEARCH SIGNIFICANCE with a guaranteed ultimate strength of 1860 MPa (270 ksi).
This research provides information on how specimen The main characteristics were those according to the manu-
cross section size-effect influences strand bond behavior and facturer: diameter, 13 mm (0.5 in.); cross-sectional area,
concrete strains at prestress transfer. This paper analyzes 100 mm2 (0.154 in.2); ultimate strength, 200.3 kN (45.1 kip);
series of tests conducted on pretensioned prestressed yield stress at 0.2%; 189.9 kN (42.8 kip); and the modulus
concrete prismatic specimens using a testing technique of elasticity, 203,350 MPa (29,500 ksi). The prestressing
based on bond behavior analysis by measuring prestressing strand was used under the as-received condition (rust-free
strand force. The tests provide data on concrete strains, and lubricant-free). The strand was not treated in any special
transfer length, effective prestressing force, bond stress, way. The strand was stored indoors and care was taken to not
and concrete modulus of elasticity. A coefficient to account drag the strand along the floor.
for the specimen cross section size-effect on the concrete
modulus of elasticity is proposed. The experimental results Testing program
have been compared with predictions from ACI 318-112 and The variables considered in the test program were
fib Model Code 2010.9 concrete mixture, specimen cross section, and concrete age
at prestress transfer. A series of specimens with different
EXPERIMENTAL RESEARCH embedment lengths were tested for each combination of
An experimental program was conducted and the ECADA variables selected. Embedment lengths followed increments
test method17 was used. The method allows for the analysis of 50 mm (2 in.) in the nearness of the transfer length for
of bond behavior by the sequential reproduction of the its determination and specimens with 1350 mm (53.15 in.)
prestress transfer and the anchorage of prestressing strands were included to determine profiles of longitudinal concrete
on the same specimen. A series of specimens with different surface strains.
embedment lengths is required to determine both transfer Specimens were designed as M-D-T-L, where M is the
and development lengths25,26 by means of the ECADA concrete mixture type (A, B, or C); D is the specimen cross
test method. Its feasibility has been verified for both short- section size in mm (100, 80, or 60, for a 100 x 100 mm2
time27,28 and long-term analyses.29,30 In this work, only [3.94 x 3.94 in.2], 80 x 80 mm2 [3.15 x 3.15 in.2], and a 60
the transfer length test results and analyses at prestress x 60 mm2 [2.36 x 2.36 in.2] cross section, respectively); T is
transfer were included. Complementarily, several speci- the concrete age at prestress transfer (12, 24, or 48 hours);
mens were also instrumented to obtain longitudinal concrete and L is the specimen embedment length (in mm).
surface strains. The designation for a complete series included only
the parameters detailing M-D-T. In two cases (A-60-48,
Materials A-80‑12), only specimens with an embedment length of
Three concrete mixtures applicable for the precast 1350 mm (53.15 in.) were made.
prestressed concrete industry with different compressive

420 ACI Structural Journal/March-April 2014


Table 2—Test program
Age Concrete (specimen embedment lengths, mm)
at prestress transfer,
hours A B C
A-80-12 (1350) — —
12
A-100-12 (300, 400, 500, 550, 600, 650, — —
700, 750, 900, 1350)
A-60-24 (300, 450, 600, 950, 1350) — —

A-80-24 (300, 350, 400, 450, 500, 550, B-80-24 (550, 600, 650, 700, 750, 800, —
24 600, 650, 700, 750, 1350) 1350)

A-100-24 (300, 400, 450, 500, 550, 600, B-100-24 (300, 550, 600, 650, 700, 750, C-100-24 (600, 650, 700, 750, 800, 1350)
650, 700, 750, 900, 1350) 800, 850, 1350)
A-60-48 (1350) B-60-48 (300, 350, 400, 450, 500, 550, —
600, 650, 950, 1000, 1350)

— B-80-48 (300, 450, 500, 550, 600, 650, C-80-48 (550, 600, 650, 800, 850, 900,
48
700, 950, 1350) 950, 1350)

— B-100-48 (300, 400, 450, 500, 550, 600, C-100-48 (550, 600, 650, 700, 750, 1350)
650, 700, 750, 800, 850, 1350)
Note: 1 mm = 0.04 in.

Fig. 3—General view of test equipment.


Fig. 2—Pretensioning frame layout.
2000 mm (78.74 in.) in length and both end frame plates
Table 2 summarizes the test program conducted and of 320 x 320 x 50 mm (12.60 x 12.60 x 1.97 in.), and an
provides the embedment lengths of tested specimens. anchorage plate of 320 x 200 x 60 mm (12.60 x 12.60 x
2.36 in.) supported by both 250 x 120 x 20 mm (9.84 x 9.84
Test equipment and instrumentation x 0.79 in.) welded elements 250 mm (9.84 in.) in length
The ECADA test method is based on measuring and (white components).
analyzing the prestressing strand force in a series of preten- The instrumentation used included a hydraulic jack pres-
sioned prestressed concrete specimens with different sure transducer to control the tensioning and detensioning
embedment lengths. Specimens are made and tested using operations; the AMA system included a hollow force trans-
pretensioning frames, as shown in Fig. 2. In this way, each ducer to measure the prestressing strand force; and in the
specimen has only one end zone with the corresponding specimens with an embedment length of 1350 mm (53.15 in.),
transfer length. detachable mechanical strain gauges were used to obtain the
Several components were coupled at both ends of the longitudinal concrete surface strain profile at the prestressing
pretensioning frames to complete the test equipment. A strand level. The gauge points were uniformly spaced at 50
hollow hydraulic jack of 300 kN (67.5 kip) capacity with mm (2 in.) intervals and were placed on two opposite spec-
an end-adjustable anchorage device was placed at one end imen faces. No internal measuring devices were used in the
to carry out tensioning, provisional anchorage, and deten- test specimens to not distort the bond phenomenon.
sioning of prestressing strands. At the opposite end, an
Anchorage-Measurement-Access (AMA) system was Test procedure
placed to guarantee the anchorage of the prestressing strand Each ECADA test series included a variable number of
and to simulate the specimen’s sectional stiffness. The AMA specimens (usually 6 to 12). The step-by-step test procedure
system was made up of a sleeve of 120 mm (4.72 in.) in for each specimen is described in detail in Martí-Vargas
length in the final specimen stretch (beyond the specimen et al.17 and may be summarized as follows (steps marked
embedment length), the end frame plate and an anchorage by “opt” have been included only for any specimens addi-
plate supported on the frame by two separators. tionally instrumented to obtain the longitudinal concrete
Figure 3 shows a general view of the test equipment, surface strains):
which includes a pretensioning frame (blue components) of a) Fabrication stage:

ACI Structural Journal/March-April 2014 421


1. The strand is placed in the frame with both anchorage
devices at the ends.
2. Strand tensioning is done using the hydraulic jack.
3. Provisional strand anchorage is done by the end-ad-
justable device.
4. The concrete is mixed, placed into the formwork in
the frame, and consolidated.
5. The specimen is cured to achieve the desired concrete
properties at the time of testing.
b) Preparation stage:
1. The formwork is removed from the frame.
2. (opt) Attaching gauge points by epoxy glue along
both the lateral sides of the specimen at the prestressing
strand level.
3. (opt) Reading the initial set of distances between Fig. 4—Transfer length determination according to ECADA
gauge points. test method.
4. The end-adjustable strand anchorage is relieved using
the hydraulic jack. when these strains were plotted according to specimen
c) Prestress transfer. Strand detensioning is produced by embedment length.
unloading the hydraulic jack. The specimen is supported on
the end frame plate. Transfer length
d) Stabilization period. The prestressing strand force Approximation to transfer length determination can be
depends on not only the strain compatibility with the obtained directly from the longitudinal concrete strain profile
concrete specimen but also on its action in the AMA system. of a specimen with a long embedment length (1350 mm
This force requires a stabilization period to guarantee its [53.15 in.] in this research), such as ascendent branch length.
measurement. In this work, the 95% Average AMS Method7 was used.
e) Measurement: On the other hand, for a complete series of specimens
1. Measuring the prestressing strand force achieved (Pi) tested with the ECADA test method, the measured Pi forces
in the AMA system. can be ordered according to specimen embedment lengths
2. (opt) Rereading the set of distances between gauge (Fig. 4). The obtained curve presented a bilinear tendency,
points (after prestress transfer). with an ascendent initial branch and horizontal branch corre-
sponding to effective prestressing force Pe. Transfer length
Test parameters Lt corresponded to the embedment length that marked the
All the specimens were prestressed by a concentrically beginning of the horizontal branch—that is, it corresponded
located single strand at a prestress level before releasing to the shorter specimen embedment length, where Pi = Pe.
75% of the guaranteed ultimate strength. Specimens were
subjected to the same consolidation and curing conditions. Effective prestressing force
The prestress transfer was gradually performed in each spec- Beyond transfer length, compatibility of the strains between
imen at a controlled speed of 0.80 kN/s (0.18 kip/s). A 2-hour the prestressing strand and the concrete exists: the prestressing
stabilization period after detensioning was established. strand strain change Δεp accounted for just before the prestress
transfer is equal to the concrete strain change Δεc. Therefore,
DATA PROVIDED FROM TESTS effective prestressing strand force can be obtained from the
The measured prestressing strand forces and sets of concrete strains according to Eq. (1)
distances between gauge points were the direct data
collected from the specimen test. Based on these direct data Pes = P0 – ΔP = P0 – Δεp · Ep · Ap (1)
from a specimen or from complete series of specimens, and
by means of back-calculations using theory of mechanics where Pes is the effective prestressing strand force obtained
concepts, parameters such as concrete strains, transfer from the measured specimen strains; P0 is the prestressing
length, effective prestressing force, bond stress, and concrete strand force just before the prestress transfer; ΔP is the
modulus of elasticity can be obtained. prestress loss due to elastic concrete shortening (ΔP = Δεp ·
Ep · Ap); Δεp is the prestressing strand strain change beyond
Concrete strains the transfer length, accounted for just before the prestress
Concrete strains can be obtained from the changes in transfer (Δεp = Δεc); Ep is the modulus of elasticity of the
distances between gauge points before and after the prestress prestressing strand; and Ap is the prestressing strand area.
transfer by dividing them by gauge length. The strain change In addition, effective prestressing strand force can be
for each gauge length was assigned to its center point. From measured directly from the AMA system on specimens with
the specimen free end, a profile with an ascendent branch, an embedment length equal to or longer than the transfer
followed by a practically horizontal branch, was depicted length. An ideal AMA system must have the same sectional
stiffness as the specimen,17 which depends on the concrete

422 ACI Structural Journal/March-April 2014


Fig. 6—Concrete strain profile for Specimen B-100-48-1350.
just before prestress transfer; n is the initial steel modular
ratio (n = Ep/Eci, where Ep is the modulus of elasticity of
the prestressing strand and Eci is the concrete modulus of
Fig. 5—End-discontinuity effect. elasticity at prestress transfer); ρ is the geometric ratio
(ρ = Ap/Ac, where Ap is the prestressing strand area and Ac is
properties and the specimen cross section. Different AMA the net cross-sectional area of the specimen).
system designs should be devised for different test condi-
tions. However, it would not be feasible to design a system EXPERIMENTAL RESULTS AND DISCUSSION
for each specific test condition. For this reason, the stiffness
of the designed AMA system was greater than the speci- Concrete strains
men’s sectional stiffness. Consequently, the prestressing By way of example based on the collected test data,
strand force measured in the AMA system after the prestress Fig. 6 provides the concrete strain profile for Specimen
transfer was greater than the effective prestressing force B-100-48-1350. The results are obtained by averaging
in the specimen. This difference of forces caused an end- the readings from the two opposite specimen faces. Three
discontinuity effect and gave rise to a slight overestimation regions can be distinguished: initial branch, plateau, and
of the real transfer length (Fig. 5). end-discontinuity. In this case, a transfer length of 600 mm
(23.6 in.) is observed directly from the curve, and another of
Bond stress 575 mm (22.6 in.) is determined from the 95% AMS.
Based on the equilibrium of the effective prestressing Beyond the transfer length, greater concrete strains—
strand force achieved, the average bond stress along the Δεci = 0.00075—result because of the early age of concrete
tranfer length can be obtained according to Eq. (2) at prestress transfer. The influence of specimen cross
section size-effect on the average concrete strains beyond
Pe the transfer length is depicted in Fig. 7. A strong influence
ut = (2)
4  is observed and explained because of the combined effects
 3 πf Lt of the different concrete stress levels and the deforma-
bility behavior related to the specimens’ cross sections.
The concrete stress level can be obtained by dividing the
where ut is the average bond stress along the transfer length; effective prestressing force, transferred according to Eq.
Pe is the effective prestressing force; (4πφ/3) is the actual (1), between the specimen’s net cross-sectional area and
seven-wire strand perimeter; φ is the nominal diameter of the concrete compressive strength at prestress transfer. As
prestressing strand; and Lt is the transfer length. Fig. 8 depicts, for the same concrete type and concrete age
at prestress transfer, the concrete stress level and concrete
Concrete modulus of elasticity strains evidently increase when the specimen cross section
An early concrete modulus of elasticity at prestress decreases. Besides, the specimens with different concrete
transfer for each specimen can be obtained from prestress stress levels present similar concrete strains when spec-
loss due to elastic concrete shortening and the transformed imens are of the same cross section size: approximately
cross-section properties according to Eq. (3) 0.0008 for specimens with a larger cross section, between
0.0012 and 0.0016 for specimens with an intermediate cross
P0 section, and approximately 0.0020 for specimens with a
− E p Ap
P0 1 ∆e ci smaller cross section. However, for the same concrete stress
∆e ci = ⋅ or Eci = (3)
1 + nρ Eci Ac Ac level, concrete strains significantly increase when the cross

section size reduces. This implies that a significant deforma-
bility behavior relating to the specimen cross section exists.
where Δεci is the elastic shortening strain of concrete due
This fact will be analyzed later by considering the concrete
to the prestress transfer; P0 is the prestressing strand force
modulus of elasticity.

ACI Structural Journal/March-April 2014 423


Fig. 9—Prestressing strand forces for Series B-100-48.
Fig. 7—Concrete strain versus specimen side cross section.

Fig. 8—Concrete strain versus concrete stress level. Fig. 10—Measured and predicted transfer length.

Transfer length 100 mm2 (3.94 x 3.94 in.2) and the 80 x 80 mm2 (3.15 x
Figure 9 provides the transferred prestressing forces 3.15 in.2) specimen cross sections. However, A-60-48 and
versus the embedment lengths for the complete B-100-48 B-60-48 show the shortest transfer lengths, while A-60-24
series. All the test specimens with embedment lengths presents a high value.
equal to or longer than 650 mm (25.6 in.) present similar Pi The transfer lengths predicted according to ACI 318-112
values and are, therefore, equal to the effective prestressing and fib Model Code 20109 from the measured parameters are
force Pe. In contrast, all the test specimens with embed- included in Fig. 10. As observed, the predictions from ACI
ment lengths shorter than 650 mm (25.6 in.) present lower 318-112 have similar values, as only strand parameters are
Pi values. Therefore, the transfer length determined by the considered. However, the predictions from fib Model Code
ECADA test method can be affirmed as 650 mm (25.6 in.) 20109 vary vastly as concrete properties are also consid-
in this case. ered. ACI 318-112 overestimates transfer length, except for
Figure 10 shows the transfer length obtained from Concrete Mixture C, while fib Model Code 20109 generally
concrete strains (directly [two values are depicted when overestimates it (only case A-60-24 is underestimated).
the ascendent branch length is unclear] and by 95% AMS)
and from prestressing strand forces. Specimens have been Effective prestressing force
ordered by concrete mixture by increasing cross section size Figure 11 shows the measured prestressing strand forces
and concrete age at prestress transfer. in the AMA system (P0, Pe), and the effective prestressing
As observed in Fig. 10, the transfer lengths in the speci- forces (Pes) obtained according to Eq. (1). For each complete
mens made with Concrete C are longer than those in the spec- series, P0 and Pe are obtained by averaging the corresponding
imens made with Concrete B which, in turn, are longer than prestressing strand forces P0 and Pi from those specimens
the transfer lengths in the specimens made with Concrete with an embedment length equal to or longer than the
A: C-100-48/B-100-48, C-100-24/B-100-24/A-100-24, transfer length. Due to the end-discontinuity effect, overesti-
C-80-48/B-80-48, B-80-24/A-80-24, and B-60-48/A-60-48. mation of the effective prestressing strand force is observed;
Generally, transfer length values also reduce when concrete Pe is always greater than Pes for all the specimens. Effec-
age at prestress transfer increases: A-60-24/A-60-48, tive prestressing force increases within the same concrete
A-80-12/A-80-24, A-100-12/A-100-24, B-80-24/B-80-48, mixture when the specimen cross section and the concrete
and C-100-24/C-100-48, except for B-100-24/B-100-48. age at prestress transfer increase. These tendencies are seen
Besides, similar transfer lengths are obtained for the 100 x more clearly by the specimen strains.

424 ACI Structural Journal/March-April 2014


Fig. 11—Effective prestressing force. Fig. 12—Average bond stress along transfer length.
Bond stress
According to Eq. (2), Fig. 12 shows the average bond
stress along the measured transfer lengths from the two
techniques, based on strains or forces. Predicted bond stress
values from ACI 318-112 and fib Model Code 20109 have
been included for comparison purposes: a constant value of
2.76 MPa (0.4 ksi)32 results from ACI 318-11,2 and 3.2fctdi
(0.464fctdi)12—where fctdi is the specified tensile strength
at prestress transfer—results from fib Model Code 2010.9
Besides, a corrected value has been obtained from fib Model
Code 20109 by applying a 0.75 factor based on the ratio
between nominal and actual strand perimeter (πφ in fib
Model Code 20109 and 4πφ/3 in ACI 318-112 and Eq. (2),
respectively).33
As observed in Fig. 12, the predictions from fib Model Fig. 13—Average bond stress versus specimen side cross
Code 20109 are of the measured values order and follow the section.
logical sequence for the different concrete types and within
cross-section size, the bond stress values rise when concrete
the same concrete type by varying specimen cross-section size
age at prestress transfer increases (except for Concrete B and
and concrete age at prestress transfer, except for A-60-24
side 100 mm [3.94 in.]).
and A-60-48. The predictions from ACI 318-112 and the
corrected predictions from fib Model Code 20109 show
Concrete modulus of elasticity: cross section
global tendencies in accordance with the transfer lengths
size-effect
depicted in Fig. 10; but inversely, however, in this case,
Figure 14 shows the experimental Eci value according
transfer length increases when bond stress decreases. As
to Eq. (3) for each specimen with an embedment length of
observed in Fig. 12, the bond stresses in the specimens made
1350 mm (53.15 in.). As seen, higher Eci values are obtained
with Concrete A are greater than those in the specimens
when concrete compressive strength increases, when spec-
made with Concrete B which, in turn, are greater than the
imen cross section increases, and when concrete age at
bond stresses in the specimens made with Concrete C.
prestress transfer increases.
The specimen cross section size-effect on the average
The early concrete modulus of elasticity is commonly
bond stresses can be analyzed from Fig. 13: a) greater bond
only obtained from the concrete compressive strength at 28
stress values for larger specimen cross sections are observed
days. To this end, as ACI 318-112 overestimates the concrete
in some cases (A-24h, B-24h, and C-48h); b) similar bond
modulus of elasticity for very early-age and concrete
stress values are obtained for A-12h irrespectively of the
compressive strengths above 50 MPa (7.25 ksi),37 the fib
specimen cross sections; and c) bond stress values decrease
Model Code 20109 provisions (Eq. (4) to Eq. (6)) are taken
for B-48h when the specimen cross section increases. Some
as a reference
authors have found these tendencies: examples of reported
increases in bond strength with increases of concrete 1/ 3
cover thickness (or specimen cross section) can be found f 
Ec 28 = 19, 400  cm  (Eq. 5.1 − 21)9 (4)
in the literature34,35 and also examples of a size effect on  10 

bond strength, which reduces as concrete cover thickness
increases.36 Finally, Fig. 13 also shows that, for the same
Ec(t) = [βcc(t)]0.5 · Ec28 (Eq. 5.1-56)9 (5)

ACI Structural Journal/March-April 2014 425


Fig. 14—Experimental concrete modulus of elasticity at Fig. 15—Comparison of concrete modulus of elasticity at
prestress transfer. prestress transfer.
   28  0.5  
bcc (t ) = exp s ⋅ 1 −     (Eq. 5.1 − 51)9 (6)
   t   

where Ec28 is the concrete modulus of elasticity at 28 days


(MPa); fcm is the concrete compressive strength at 28 days
(MPa); Ec(t) is the concrete modulus of elasticity at age t;
βcc(t) is a coefficient to describe development with time;
s is a coefficient to account for the strength class of cement
(s = 0.2 for Class 52.5 R); and t is concrete age (days).
To analyze the specimen cross section size-effect on the
concrete modulus of elasticity from the obtained test results,
several adjustments of the experimental Eci values (Eq. (3))
and the expected Ec(t) values at prestress transfer (Eq. (5))
are made as shown in Fig. 15. Therefore, the following equa- Fig. 16—Comparison of concrete modulus of elasticity at 28
tion is proposed days.
The results of this process are presented in Table 3.
Eci = λ · Ec(t) (7) Also, the Ec28_s values can be obtained approximately
from the concrete compressive strength through Ec28 (Eq.
where λ is a coefficient to account for the cross-section (4)) using Eq. (7): Ec28_s = λ· Ec28; and an expected Eci can
size-effect: 0.721 for 100 x 100 mm2 (3.94 x 3.94 in.2) spec- be computed using Eq. (5), by substituting Ec(t) for Eci and
imen cross section, 0.612 for 80 x 80 mm2 (3.15 x 3.15 in.2), Ec28 for Ec28_s.
and 0.523 for 60 x 60 mm2 (2.36 x 2.36 in.2).
The concrete modulus of elasticity for each specimen at Concrete modulus of elasticity: influence of
28 days (Ec28_s) can be obtained as follows: concrete stress
1. Determination of Eci (Eq. (3)). A normalized concrete modulus of elasticity Eci* = Eci/λ
2. Computation of a concrete modulus of elasticity at 28 (Eci from Eq. (3) and λ from Eq. (7)) is obtained to avoid
days (Ec28_s) using Eq. (5) with Ec(t) = Eci. the cross section size-effect. Figure 17 depicts the obtained
3. Application of the λ coefficient (Eq. (7)), as follows: Eci* relating to the concrete stress level. As observed, for the
Ec28_s = λ· Ec28_ref. same concrete type and concrete age at prestress transfer,
4. The Ec28_ref values are obtained. These values have the tendencies when the concrete stress level increases are
good agreement with the fib Model Code 20109 provisions as follows: Eci* decreases for Series A-12 h and B-24 h and
(Eq. (4)) for each concrete mixture, as shown in Fig. 16. Eci* increases for Series C-48 h, with Eci* practically nothing
Therefore, the λ coefficient—initially obtained at prestress varies for Series A-24 h and B-48 h. Besides, a general
transfer—is also applicable at 28 days. tendency showing similar Eci* values for different concrete
5. An Ec28_m value is computed for each concrete mixture stress levels (refer to the horizontal line) is observed when
by averaging the Ec28_ref values from the specimens made considering all the series. Therefore, the concrete modulus
with the same concrete mixture. of elasticity seems to be practically independent of concrete
6. The Ec28_s values are obtained using Eq. (7): Ec28_s = λ· stress, and the main influence is based on the specimen cross
Ec28_m. section size-effect.

426 ACI Structural Journal/March-April 2014


Table 3—Concrete modulus of elasticity, MPa
Modulus of elasticity of concrete, MPa
Experimental fib Model Code 20109
Specimen Eci Ec28_s Ec28_ref Ec28_m Ec28_s Ec28 Ec28_s Eci
A-60-24 13,790 21,180 40,498 20,848 20,707 13,481
A-60-48 16,205 21,317 40,759 20,848 20,707 15,741
A-80-12 11,221 21,458 35,062 24,396 24,230 12,670
39,863 39,592
A-80-24 15,820 24,299 39,703 24,396 24,230 15,775
A-100-12 16,718 31,971 44,342 28,741 28,546 14,927
A-100-24 18,219 27,984 38,812 28,741 28,546 18,585
B-60-48 12,296 16,175 30,927 17,190 17,350 13,189
B-80-24 13,146 20,191 32,992 20,116 20,302 13,218
B-80-48 16,037 21,095 34,470 32,868 20,116 33,174 20,302 15,434
B-100-24 16,002 24,578 34,088 23,698 23,918 15,572
B-100-48 17,465 22,975 31,865 23,698 23,918 18,183
C-80-48 15,366 20,213 33,027 19,320 19,307 14,677
C-100-24 14,742 22,643 31,405 31,569 22,761 31,547 22,745 14,809
C-100-48 16,594 21,828 30,275 22,761 22,745 17,291

Note: 1 MPa = 145 psi.

Fig. 17—Concrete modulus of elasticity versus concrete


stress level. Fig. 18—Comparison of elastic shortening losses.

Concrete modulus of elasticity: elastic shortening for those specimens with a smaller cross section. This fact
prestress loss can be explained by the different concrete stress levels and
Finally, to verify the applicability of the obtained λ coef- the deformability behavior relating to cross sections of spec-
ficients to account for the cross section size-effect, predicted imens. Therefore, a clear effect of specimen cross section
prestress losses due to the elastic shortening of concrete size on prestress loss cannot be ruled out.
from the measured and theoretical parameters are estimated Moreover, Fig. 18 shows that the tendencies of the
using preexisting equations. According to the ACI 318-11 measured prestress losses according to the variable’s
Commentary,2 prestress losses can be calculated in accor- concrete mixture, specimen cross section size, and concrete
dance with several procedures.38,39 Figure 18 shows the age at prestress transfer are followed by the prestress losses
measured and predicted prestress losses due to the elastic predicted by all the methods: prestress losses disminish in
shortening of concrete for all the specimens with an embed- the same concrete mixture when the specimen cross section
ment length of 1350 mm (53.15 in.). The λ coefficient has increases and when the concrete age at prestress transfer
been applied in all cases. increases, and prestress losses in the specimens made with
As observed in Fig. 18, the predicted prestress losses Concrete C are greater than those in the specimens made
follow the tendencies of measured prestress losses. The with Concrete B, which are also greater than the prestress
prestress loss ranges with values of 10% for those specimens losses in the specimens made with Concrete A. The PCI
with a larger cross section, with values of 15 to 20% for CPL39 predictions made with experimental Eci (Eq. (3)) and
those with an intermediate cross section, and of 25 to 30% fib Model Code 20109 practically coincide with the measured

ACI Structural Journal/March-April 2014 427


prestress losses. However, the PCI CPL39 predictions made elements, fiber-reinforced concrete, durability of concrete structures, and
strut-and-tie models.
with Eci according to ACI 318-112 underestimate prestress
losses as the concrete modulus of elasticity is overestimated. Libardo A. Caro is an Assistant Researcher and PhD Candidate in the
For Zia et al.38 predictions, which consider gross section Department of Construction Engineering and Civil Engineering Projects at
UPV. He received his civil engineering degree from the Universidad Santo
properties, an overestimation trend of prestress losses for Tomás, Bogotá, Colombia. His research interests include bond properties
cases with smaller specimen cross section sizes is observed. of prestressed concrete structures and the use of advanced cement-based
materials in structural applications.
CONCLUSIONS Pedro Serna-Ros is a Professor of civil engineering at UPV, where he
Based on the results of this experimental research, the received his MEng in civil engineering; he received his PhD from École
following main conclusions can be drawn: Nationale des Ponts et Chaussées, Paris, France. His research interests
include self-consolidating concrete, fiber-reinforced concrete, and the bond
• Transfer lengths have been determined by two tech- behavior of reinforced and prestressed concrete.
niques: the longitudinal concrete strain profile and
prestressing strand forces. Based on both techniques, ACKNOWLEDGMENTS
transfer length values are higher when concrete Funding for this research has been provided by the Spanish Ministry of
Education and Science, the Ministry of Science and Innovation, and ERDF
quality diminishes. (Projects BIA2006-05521 and BIA2009-12722).
• The transfer lengths predicted from ACI 318-112 have
similar values, irrespectively of concrete quality, as REFERENCES
only strand parameters are considered. However, the 1. Janney, J., “Nature of Bond in Pretensioned Prestressed Concrete,”
fib Model Code 20109 predictions vary considerably as ACI Journal, V. 51, No. 5, May 1954, pp. 717-737.
2. ACI Committee 318, “Building Code Requirements for Structural
concrete properties are also considered. ACI 318-112 Concrete (ACI 318-11) and Commentary,” American Concrete Institute,
overestimates transfer length when concrete quality Farmington Hills, MI, 2011, 473 pp.
is good, while fib Model Code 2010 overestimates 3. fib, “Bond of Reinforcement in Concrete. State-of-Art Report,”
Fédération Internationale du Béton, Bulletin d’Information No. 10, Laus-
in general. anne, Switzerland, 2000, 427 pp.
• A strong influence of specimen cross section size-ef- 4. ElBatanouny, M. K.; Ziehl, P. H.; Larosche, A.; Mays, T.; and
fect on average concrete strains beyond transfer length Caicedo, J. M., “Bent-Cap Confining Stress Effect on the Slip of
Prestressing Strands,” ACI Structural Journal, V. 109, No. 4, July-Aug.
has been observed. This fact can be explained by the 2012, pp. 487-496.
combined effects of different concrete stress levels and 5. ElBatanouny, M. K., and Ziehl, P. H., “Determining Slipping Stress of
deformability behavior in relation to the cross sections Prestressing Strands in Confined Sections,” ACI Structural Journal, V. 109,
No. 6, Nov.-Dec. 2012, pp. 767-776.
of specimens. 6. Barnes, R. W.; Grove, J. W.; and Burns, N. H., “Experimental Assess-
• The effective prestressing force increases in the same ment of Factors Affecting Transfer Length,” ACI Structural Journal, V.
concrete mixture when specimen cross section and 100, No. 6, Nov.-Dec. 2003, pp. 740-748.
7. Russell, B. W., and Burns, N. H., “Measured Transfer Lengths of 0.5
concrete age at prestress transfer increase. These tenden- and 0.6 in. Strands in Pretensioned Concrete,” PCI Journal, V. 44, No. 5,
cies are seen more clearly with the specimen strains Sept.-Oct. 1996, pp. 44-65.
technique rather than with the strand forces technique. 8. CEN, “Eurocode 2: Design of Concrete Structures—Part 1-1: General
Rules and Rules for Buildings,” European Standard EN 1992-1-1:2004:E,
• Higher bond stress values for larger specimen cross Comité Européen de Normalisation, Brussels, Belgium, 2004, 225 pp.
sections have been observed in most cases. For the same 9. fib, “Model Code 2010. First complete draft—Volume 1,” Fib Bulletin
cross section size, bond stress values are greater when No. 55, International Federation for Structural Concrete, Lausanne, Swit-
zerland, 2010, 292 pp.
the concrete age at prestress transfer increases and when 10. AENOR, “UNE 36094: Alambres y Cordones de Acero para Arma-
concrete quality increases. duras de Hormigón Pretensado,” Asociación Española de Normalización y
• The concrete modulus of elasticity at prestress transfer Certificación, Madrid, Spain, 1997, 21 pp.
11. ASTM A416/A416M-10, “Standard Specification for Steel Strand,
has been obtained from the experimental data of Uncoated Seven-Wire for Prestressed Concrete,” ASTM International,
prestressed specimens. Higher concrete modulus of West Conshohocken, PA, 2010, 5 pp.
elasticity values result from greater concrete compres- 12. Martí-Vargas, J. R.; Arbeláez, C. A.; Serna-Ros, P.; Navarro-Gregori,
J.; and Pallarés-Rubio, L., “Analytical Model for Transfer Length Predic-
sive strength, larger specimen cross section, and older tion of 13 mm Prestressing Strand,” Structural Engineering & Mechanics,
concrete age at prestress transfer. V. 26, No. 2, 2007, pp. 211-229.
• A coefficient to account for the specimen cross section 13. Martí-Vargas, J. R.; Serna, P.; Navarro-Gregori, J.; and Pallarés,
L., “Bond of 13 mm Prestressing Steel Strands in Pretensioned Concrete
size-effect on the concrete modulus of elasticity Members,” Engineering Structures, V. 41, 2012, pp. 403-412.
is proposed. 14. Moustafa, S., “Pull-out Strength of Strand and Lifting Loops,”
• fib Model Code 2010 predictions for transfer length, Technical Bulletin 74-B5, Concrete Technology Associates, Tacoma, WA,
1974, 34 pp.
average bond stress, concrete modulus of elasticity at 15. Cousins, T. E.; Badeaux, M. H.; and Moustafa, S., “Proposed Test
prestress transfer, and elastic shortening prestress loss for Determining Bond Characteristics of Prestressing Strand,” PCI Journal,
have showed a best agreement to the experimental V. 37, No. 1, Jan.-Feb. 1992, pp. 66-73.
16. Peterman, R. J., “A Simple Quality Assurance Test for Strand Bond,”
results than the ACI 318-112 predictions. PCI Journal, V. 54, No. 2, 2009, pp. 143-161.
17. Martí-Vargas, J. R.; Serna-Ros, P.; Fernández-Prada, M. A.; Miguel-
AUTHOR BIOS Sosa, P. F.; and Arbeláez, C. A., “Test Method for Determination of the
José R. Martí-Vargas is an Associate Professor of civil engineering at Transmission and Anchorage Lengths in Prestressed Reinforcement,”
the Universitat Politècnica de València (UPV), València, Spain, where he Magazine of Concrete Research, V. 58, No. 1, Feb. 2006, pp. 21-29.
received his MEng in civil engineering and his PhD. His research interests 18. Thorsen, N., “Use of Large Tendons in Pretensioned Concrete,” ACI
include the bond behavior of reinforced and prestressed concrete structural Journal, V. 52, No. 2, Feb. 1956, pp. 649-659.

428 ACI Structural Journal/March-April 2014


19. Guyon, Y., “Béton Précontrainte. Étude Théorique et Expérimen- 30. Caro, L. A.; Martí-Vargas, J. R.; and Serna, P., “Time-Dependent
tale,” Ed. Eyrolles, Paris, France, 1953, 711 pp. Evolution of Strand Transfer Length in Pretensioned Prestressed Concrete
20. Balázs, G., “Transfer Length of Prestressing Strand as a Function of Members,” Mechanics of Time-Dependent Materials, V. 17, No. 4, Nov.
Draw-in and Inicial Prestress,” PCI Journal, V. 38, No. 2, Mar.-Apr. 1993, 2013, pp. 501-527.
pp. 86-93. 31. EN 197-1:2000, “Cement. Part 1: Compositions, Specifications and
21. Martí-Vargas, J. R.; Arbeláez, C. A.; Serna-Ros, P.; and Castro-Bu- Conformity Criteria for Common Cements,” Comité Européen de Normal-
gallo, C., “Reliability of Transfer Length Estimation from Strand End Slip,” isation, Brussels, Belgium, 2000, 30 pp.
ACI Structural Journal, V. 104, No. 4, July-Aug. 2007, pp. 487-494. 32. Tabatabai, H., and Dickson, T., “The History of the Prestressing
22. Mahmoud, Z. I.; Rizkalla, S. H.; and Zaghloul, E. R., “Transfer Strand Development Length Equation,” PCI Journal, V. 38, No. 5,
and Development Lengths of Carbon Fiber Reinforcement Polymers Sept.-Oct. 1993, pp. 64-75.
Prestressing Reinforcement,” ACI Structural Journal, V. 96, No. 4, 33. Martí-Vargas, J. R., and Hale, W. M., “Predicting Strand Transfer
July-Aug. 1999, pp. 594-602. Length in Pretensioned Concrete: Eurocode versus North American Prac-
23. Caro, L. A.; Martí-Vargas, J. R.; and Serna, P., “Prestress Losses tice,” Journal of Bridge Engineering, ASCE, V. 18, No. 12, Dec. 2013, pp.
Evaluation in Prestressed Concrete Prismatic Specimens,” Engineering 1270-1280.
Structures, V. 48, 2013, pp. 704-715. 34. Yerlici, V. A., and Özturan, T., “Factors Affecting Anchorage Bond
24. Deatherage, J. H.; Burdette, E.; and Chew, C. K., “Development Strength in High-Performance Concrete,” ACI Structural Journal, V. 97,
Length and Lateral Spacing Requirements of Prestressing Strand for No. 3, May-June 2000, pp. 499-507.
Prestressed Concrete Bridge Girders,” PCI Journal, V. 39, No. 1, Jan.-Feb. 35. García-Taengua, E.; Martí-Vargas, J. R.; and Serna-Ros, P., “Statis-
1994, pp. 70-83. tical Approach to Effect of Factors Involved in Bond Performance of
25. Martí-Vargas, J. R.; Arbeláez, C. A.; Serna-Ros, P.; Fernández-Prada, Steel Fiber-Reinforced Concrete,” ACI Structural Journal, V. 108, No. 4,
M. A.; and Miguel-Sosa, P. F., “Transfer and Development Lengths of July-Aug. 2011, pp. 461-468.
Concentrically Prestressed Concrete,” PCI Journal, V. 51, No. 5, Sept.-Oct. 36. Ichinose, T.; Kanayama, Y.; Inoue, Y.; and Bolander, J. E., “Size
2006, pp. 74-85. Effect on Bond Strength of Deformed Bars,” Construction & Building
26. Martí-Vargas, J. R.; Serna, P.; and Hale, W. M., “Strand Bond Perfor- Materials, V. 18, 2004, pp. 549-558.
mance in Prestressed Concrete Accounting for Bond Slip,” Engineering 37. Khan, A. A.; Cook, W. D.; and Mitchell, D., “Early Age Compressive
Structures, V. 51, 2013, pp. 236-244. Stress-Strain Properties of Low-, Medium-, and High-Strength Concretes,”
27. Martí-Vargas, J. R.; Serna-Ros, P.; Arbeláez, C. A.; and Riguei- ACI Materials Journal, V. 92, No. 6, Nov.-Dec. 1995, pp. 617-624.
ra-Victor, J. W., “Bond Behaviour of Self-Compacting Concrete in Trans- 38. Zia, P.; Preston, H. K.; Scott, N. L.; and Workman, E. B., “Esti-
mission and Anchorage,” Materiales de Construcción, V. 56, No. 284, mating Prestress Losses,” Concrete International, V. 1, No. 6, June 1979,
2006, pp. 27-42. pp. 32-38.
28. Martí-Vargas, J. R.; Serna, P.; Navarro-Gregori, J.; and Bonet, J. L., 39. PCI Commitee on Prestress Losses, “Recommendations for Esti-
“Effects of Concrete Composition on Transmission Length of Prestressing mating Prestress Losses,” PCI Journal, V. 20, No. 4, July-Aug. 1975,
Strands,” Construction & Building Materials, V. 27, 2012, pp. 350-356. pp. 43-75.
29. Martí-Vargas, J. R.; Caro, L.; and Serna, P., “Experimental Technique
for Measuring the Long-Term Transfer Length in Prestressed Concrete,”
Strain, V. 49, 2013, pp. 125-134.

ACI Structural Journal/March-April 2014 429


NOTES:

430 ACI Structural Journal/March-April 2014


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 111-S38

Proposed Minimum Steel Provisions for Prestressed and


Nonprestressed Reinforced Sections
by Natassia R. Brenkus and H. R. Hamilton
The current ACI code includes two separate provisions for minimum Explicit consideration of the cracking moment when
steel reinforcement: one for nonprestressed reinforced sections, determining minimum reinforcement leads to an iterative
and another for prestressed sections. For nonprestressed reinforced problem, sometimes without solution. Prestressed concrete
sections, the current minimum steel requirement is written in terms only cracks when the applied flexural tensile stress exceeds
of geometric and material properties of the section. For prestressed
both the net compressive stress from the prestressing force
concrete, the current provision is written in terms of the cracking
in the steel and the tensile strength of the concrete. Conse-
moment of the section. Prestressed concrete only cracks when the
applied flexural tensile stress exceeds both the tensile strength of the quently, when bonded prestressing steel quantities are
concrete and the net compressive stress from the prestressing force increased, the cracking moment of the section increases.
in the steel. Consequently, when bonded prestressing steel quanti- In certain instances, depending on the shape of the cross
ties are increased, the cracking moment of the section increases. section and ultimate strength requirements, it is possible for
Depending on the shape of the cross section and ultimate strength a section to contain a large volume of bonded prestressing
requirements, in certain instances it is possible that a section can steel yet fail to meet the minimum reinforcement require-
contain a large volume of bonded prestressing steel and yet not ment. This issue has been raised by others (Kleymann et al.
meet the minimum reinforcement requirement. A parametric study 2006; Freyermuth and Aalami 1997; Ghosh 1986).
of several cross sections was performed to investigate this behavior. To address this issue and the inherent problem of sepa-
This paper describes an exact and unified solution approach for
rate provisions for prestressed and nonprestressed concrete,
specifying minimum reinforcement for both nonprestressed and
a unified approach for determining minimum reinforce-
prestressed sections. A second parametric study to validate the
proposed minimum steel provisions is also presented. ment for both nonprestressed and prestressed sections was
derived, and is presented in this paper. The proposed provi-
Keywords: ductile design; flexure; minimum reinforcement; prestressed sions are based on minimum steel concepts for prestressed
concrete. concrete introduced by Leonhardt (1964), avoid the explicit
consideration of cracking moment, and are applicable to
INTRODUCTION both prestressed and nonprestressed concrete. The para-
Reinforced and prestressed concrete members are typi- metric studies presented in this paper compare the proposed
cally designed with a minimum quantity of flexural rein- provisions with current provisions and detailed calculations
forcement. Steel reinforcement ensures that the post- of moment capacity using a direct tensile concrete behavior
cracking flexural capacity of the section is greater than the approach rather than the modulus of rupture. Simplified
cracking moment Mcr, which is typically defined as the forms of the provisions are presented for rectangular and
moment necessary to cause the maximum flexural tensile T-shaped members.
stress to exceed the modulus of rupture of the concrete. If a
beam with insufficient reinforcement cracks, then the rein- RESEARCH SIGNIFICANCE
forcement will rupture immediately following crack forma- ACI 318-11 has different minimum reinforcement
tion; the resulting failure mode, being sudden and brittle, is requirements for prestressed and nonprestressed reinforce-
undesirable and dangerous. Consequently, minimum rein- ment. Furthermore, the current method of calculating the
forcement requirements have life-safety implications. minimum steel requirements for prestressed members
Both AASHTO LRFD (AASHTO 2007) and ACI 318-11 results in a variable quantity of minimum reinforcement,
(ACI Committee 318 2011) design specifications have which depends on the prestress level. Such an approach can
minimum reinforcement requirements. ACI 318-11 has sepa- result in an inefficient use of reinforcing steel or even, in
rate provisions for nonprestressed and prestressed concrete. some circumstances, the inability to satisfy the minimum
The current nonprestressed provisions do not explicitly reinforcement requirements. The derivation and parametric
consider cracking moment, while the prestressed provisions studies described in this paper provide a direct and unified
require the direct calculation of cracking moment. This is method of determining minimum steel requirements for both
true in the AASHTO LRFD as well. The AASHTO LRFD, nonprestressed reinforced and prestressed sections.
however, applies to both prestressed and nonprestressed
concrete. Although both design specifications aim to ensure
ACI Structural Journal, V. 111, No. 2, March-April 2014.
that flexural members contain a sufficient quantity of rein- MS No. S-2012-175.R1, doi:10.14359.51686531, was received July 17, 2012, and
forcement so that the post-cracking flexural capacity is at reviewed under Institute publication policies. Copyright © 2014, American Concrete
Institute. All rights reserved, including the making of copies unless permission is
least equal to the cracking moment of the concrete section, obtained from the copyright proprietors. Pertinent discussion including author’s
closure, if any, will be published ten months from this journal’s date if the discussion
the method of implementation has been somewhat different. is received within four months of the paper’s print publication.

ACI Structural Journal/March-April 2014 431


CURRENT PROVISIONS: HISTORICAL AASHTO LRFD PROVISIONS
PERSPECTIVE AASHTO LRFD specifications (2007) have unified provi-
The earliest known minimum reinforcement requirement sions that apply to sections containing either prestressed
appeared in a 1936 ACI committee document. A minimum or nonprestressed reinforcement, or both. To meet the
area of steel satisfying Eq. (1) was required for nonpre- minimum reinforcement requirements, the flexural resis-
stressed concrete (ACI Committee 501 1936) tance Mr should satisfy Eq. (3) or (4)

As,min = 0.0005b′d (1) Mr ≥ 1.2Mcr (3)

where b′ is the width of web in I- or T-beam sections, and d where Mcr is the cracking moment
is the depth from the compression face to the center of the
longitudinal tensile reinforcement. Mr ≥ 1.33Mu (4)
In 1954, the Bureau of Public Roads issued design
requirements for prestressed concrete bridges (Bureau of where Mu is the factored moment of the applicable load
Public Roads 1954) and, in 1958, general recommenda- combination.
tions were published for the design of prestressed concrete The cracking moment is determined using Eq. (5)
(ACI-ASCE Committee 323 2004). Neither of these docu-
ments, however, contained minimum reinforcement require- S 
ments for prestressed concrete. M cr = Sc ( fr + fcpe ) − M dnc  c − 1 ≥ Sc fr (5)
 nc 
S
In 1963, prestressed concrete was incorporated within
the scope of ACI 318-63 (ACI Committee 318 1963);
the committee reasoned that combined discussion of the where fr is the modulus of rupture of the concrete; fcpe is the
nonprestressed and prestressed reinforced concrete was less compressive stress in the concrete due to effective prestress
confusing than their separate consideration. To this end, ACI forces (after allowance for all prestress losses) at the extreme
318-63 combined nonprestressed reinforced and prestressed fiber of the section where tensile stress is caused by exter-
concrete under the generic term “reinforced concrete”; it nally applied loads; Sc is the section modulus with respect
included, however, two separate requirements for minimum to the extreme tensile fiber of the composite section; Snc is
tensile reinforcement. the section modulus with respect to the extreme tensile fiber
During this code cycle, the minimum reinforcement provi- of the monolithic or noncomposite section; and Mdnc is the
sion for nonprestressed concrete was rewritten in terms of p, total unfactored dead load moment acting on the monolithic
the ratio of the area of the tension reinforcement to the effec- or noncomposite section. When monolithic or noncomposite
tive area of the concrete. As a lower limit, an area of steel sections are designed to resist all loads, the designer is
satisfying 200bd/fy was required; this absolute minimum directed to substitute Snc for Sc in Eq. (5) for the calculation
was derived by equating the ultimate strength of the section of Mcr.
without reinforcement to the ultimate strength of the section The modulus of rupture fr used depends on the limit state
with reinforcement and solving for p (ACI Committee 318 being checked and the specified concrete strength. Litera-
1963). In cases where the provided area of reinforcement ture is cited in which the modulus of rupture values typically
was one-third greater than that required by analysis, the range between 7.5√fc′ psi (0.62√fc′ MPa) and 11.7√fc′ psi
minimum reinforcement requirement was considered satis- (0.97√fc′ MPa) (ACI Committee 318 1992; Walker and
fied—an exception included to ensure that the minimum Bloem 1960; Khan et al. 1996). For determining minimum
reinforcement required for large members was not excessive. steel requirements using the AASHTO LRFD provisions, the
The separate prestressed concrete design chapter added cracking moment is calculated using an estimated modulus
to the ACI code during this cycle introduced minimum of rupture equal to 11.7√fc′ psi (0.97√fc′ MPa). The AASHTO
reinforcement requirements for prestressed concrete: the LRFD rationale for using a higher modulus of rupture value
provided area of steel was required to be adequate to develop for minimum steel requirements is that it is a strength limit
an ultimate load capacity greater than 1.2 times the cracking state, so the use of the upper bound value is justified; the
load. The cracking load was based on a modulus of rupture 20% margin provided by Eq. (3) could be lost by using a
of 7.5√fc′ psi (0.62√fc′ MPa). lower modulus-of-rupture value. These provisions are valid
In 1995, the minimum reinforcement requirements for for specified concrete strengths up to 15,000 psi (100 MPa).
nonprestressed concrete were revised to explicitly account
for concrete and reinforcement tensile strengths (ACI ACI PROVISIONS
Committee 318 1995). ACI 318-11, Section 18.8.2, specifies the following for
Equation (2) gives the typical requirement minimum reinforcement:
“Total amount of prestressed and nonprestressed rein-
3 fc′ forcement in members with bonded prestressed reinforce-
As,min = bw d ≥ 200bw d / f y (2) ment shall be adequate to develop a factored load at least
fy
1.2 times the cracking load computed on the basis of the
modulus of rupture fr specified in 9.5.2.3.”

432 ACI Structural Journal/March-April 2014


The cracking load, however, is not explicitly defined in
ACI 318-11 as it is in AASHTO LRFD. In general, it is
implicitly understood that the cracking load is the force
required such that the stress in the extreme tension fiber
is equal to the modulus of rupture, which is defined as
7.5√fc′ psi (0.62√fc′ MPa).

EFFECT OF PRESTRESSING ON MINIMUM STEEL


REQUIREMENTS
As described previously, the current AASHTO LRFD
and ACI 318-11 minimum steel requirements are based
on cracking moment, either implicitly or explicitly. The
consequence of this dependence is that the quantity of
minimum reinforcement for a given cross section varies
with the prestressing force, and thus, the quantity of bonded
prestressing steel.
Oladapo (1968) noted this peculiarity in a parametric
study that compared the moment capacity for varying levels Fig. 1—Results from parametric study of current ACI provi-
of prestress, eccentricity, and tensile strength of the concrete. sions (O’Neill and Hamilton 2009).
He found that, after a certain level of prestress, continuing to where the difference between the two curves for SB in nega-
increase the amount of prestressing steel moved the section tive bending is much greater than that between the hollow
further away from meeting the minimum requirements. core curves or the Florida bulb-tee curves. Consequently,
O’Neill and Hamilton (2009) conducted a similar para- as the amount of prestressing steel increases, the cracking
metric study on a Florida bulb tee 78, a 4 ft (1.22 m) wide moment increases at a faster rate than the moment capacity
by 8 in. (20.3 cm) deep hollow core slab, and a segmental due to the overriding effect of the relatively large tension
box girder, verifying Oladapo’s findings. Figure 1 shows zone in sections such as the SB.
the results of the study. The variation of the design moment The use of 7.5√fc′ psi (0.62√fc′ MPa) versus 11.7√fc′
capacity φMn of each section was calculated using strain psi (0.97√fc′ MPa) was also investigated by O’Neill and
compatibility and normalized by the cracking moment; Hamilton (2009) in a series of prestressed girder and pile
cracking moment was calculated using Eq. (5). The moment laboratory tests. It was found that the girder cracking stresses
capacity varies as a result of the increase in area of bonded ranged from approximately 6√fc′ psi (0.5√fc′ MPa) to 14√fc′
prestressing steel Aps from zero to an arbitrary maximum; psi (1.2√fc′ MPa). In cases with low amounts of prestressing
Aps was normalized by the area of the concrete cross section steel, however, the cracking stress ranged from 6.1√fc′ psi
between the flexural tension face and the center of gravity of (0.5√fc′ MPa) to 7.6√fc′ psi (0.63√fc′ MPa). As the minimum
the gross section Act. The Florida bulb tee and hollow core steel requirement usually controls the steel quantity when the
were investigated for positive bending, and the segmental section is larger than required for strength, and the amount
box girder was investigated for negative bending. The graph of prestressing steel required is similarly small, the use of
contains two curves for each section, which correspond to 7.5√fc′ psi (0.62√fc′ MPa) as an estimate of the modulus
the ACI 318-11 modulus of rupture estimate of 7.5√fc′ psi of rupture is justified. Minimum reinforcement require-
(0.62√fc′ MPa) and AASHTO LRFD estimate of 11.7√fc′ psi ments for nonprestressed sections do not change with the
(0.97√fc′ MPa). selected quantity of reinforcement. The authors believe that
All three sections exhibit a similar trend: as bonded a more rational approach would be for nonprestressed and
prestressing steel is added, the curves quickly rise above prestressed concrete to have a single unique minimum rein-
the minimum steel required to reach 1.2Mcr. Because the forcement requirement for each section. This requires direct
cracking moment increases at a slightly greater rate than calculation of the cracking moment using only section and
that of the moment capacity, each curve eventually declines material properties, as will be shown in the following section
as more prestressing steel is added. In general, the large with the derivation of a minimum reinforcement equation.
quantities of prestressing steel required to cause a decrease
in the moment ratio are likely beyond typical strength and PROPOSED MINIMUM REINFORCEMENT
serviceability needs. In the case of the box girder in nega- PROVISIONS
tive bending, however, the addition of bonded prestressing Leonhardt (1964) proposed a minimum steel require-
steel of approximately 1.4% beyond an Aps/Act ratio results ment for prestressed concrete based on providing sufficient
in a section that does not meet AASHTO LRFD minimum steel area to resist net tensile concrete stresses that occur
reinforcement requirements. Beyond this point, unnecessary just before cracking. His approach takes advantage of the
bonded prestressing steel should be added to satisfy Eq. (3). prestressing steel’s tensile capacity between the effective
The precompression force in a section has less effect than prestressed state (State I) and the ultimate state at flexural
the concrete tensile strength on the section’s overall strength capacity (State II) without explicit consideration of the
when the compression zone is smaller than tension zone, as cracking moment to determine the minimum prestressing
in the case of the box girder (SB). This is seen in the plot steel required. It is applicable to both nonprestressed and

ACI Structural Journal/March-April 2014 433


Fig. 2—Concrete and reinforcement forces at State I and
State II (Leonhardt 1964).
Fig. 3—Stress state of section at decompression.
prestressed concrete. A representation of the concept is
shown in Fig. 2.
The method ensures sufficient area of reinforcement such
that the tensile force associated with the incremental change
in steel stress from the effective prestress fse to the stress
associated with ultimate flexural capacity fps, combined with
the nonprestressed reinforcement yield strength fy, is greater
than the tensile force generated in the concrete just before
cracking Tcr. This relationship can be expressed as

Apsfps ≥ Tcr + Apsfse + Asfy (6)

Leonhardt’s proposed method made several assumptions:


the tensile zone is rectangular, the resultant of the tension
block is at the same elevation as the prestressing steel, and
the section is noncomposite. These assumptions are prob-
ably conservative in most cases; however, to accurately
model real sections without requiring excessive steel area, it
is important to account for the variation of the steel location, Fig. 4—External and internal forces diagram.
the section shape, and the effects of composite construction.
The following derivation takes all of these variations into achieve zero stress at the extreme tensile fiber is differenti-
account by resolving the forces into the internal and external ated herein from the moment subsequently applied to achieve
moments acting on a section. To illustrate the proposed the rupture stress at the extreme tensile fiber. The cracking
method, consider a rectangular section with nonprestressed moment Mcr is then defined as the sum of the decompression
steel and prestressing steel reinforcement as shown in Fig. 3. and net cracking moment, Mdec + Mncr, and is used to deter-
The left stress profile shows the section under effective mine the minimum quantity of prestressing steel necessary
prestress after time-dependent losses have occurred; fse is the to ensure a ductile failure mode. This difference in definition
effective prestress in the tendon, and fpe is the compressive is illustrated in Fig. 3. Leonhardt (1964) reasoned (reframed
stress in concrete at the extreme fiber of section where tensile herein in terms of moment rather than force) that because
stress is caused by externally applied loads. The center the prestressing steel necessary to resist in Mdec is already
stress profile shows an externally applied moment causing present, then additional reinforcement or prestressing steel is
the concrete tensile stress at the bottom of the section to be needed only to ensure minimum strength to support Mncr. To
exactly equal in magnitude to fpe. Superimposing the two ensure that the structure remains stable after cracking, suffi-
stress profiles results in a net stress at the extreme fiber of cient reinforcement should be present such that the moment
zero. The applied moment required to cause this stress state capacity is greater than the cumulative applied moment
is defined as the decompression moment Mdec. While this Mdec + Mncr ≤ Mn. Figure 4 shows these external and internal
moment will cause an increase in prestressing steel stress, moments acting on a rectangular section.
in typical prestressed concrete members, this increase is not The left side of Fig. 4 illustrates the stress distribution in
considered significant. a concrete section just before cracking. If the stress at the
As moment is applied beyond the decompression moment, extreme fiber of the concrete is equal to the rupture strength,
the net stress in the bottom of the section becomes tensile. then the net cracking moment can be defined as
The point at which the tensile stress at the extreme fiber is
equal to the modulus of rupture defines cracking; this applied fr I
M ncr = (7)
moment is shown as the right stress profile and is referred to yt
as Mncr, or the net cracking moment. The moment applied to

434 ACI Structural Journal/March-April 2014


which ACI currently employs as an estimate of concrete
tensile strength when determining minimum reinforcement.
Further, as discussed previously, additional investigations
by O’Neill and Hamilton (2009) justify the use of this esti-
mate for the calculation of minimum steel. The net cracking
moment described in Eq. (7) can then be written

I
M ncr = 7.5 fc′ (8)
yt

Fig. 5—Stress states under decompression moment.
The force coupled on the right side of the figure represents
the internal forces immediately after cracking, in which the 7.5 fc′I  I  1 yt 
reinforcement resists the tensile force carried by the concrete fAps f ps jd p + fjdAs f y ≥ 1.2 + Aps fse   + e I  (12)
yt  Ayt  final  A transfer
before cracking. If the reinforcement volume is sufficiently
low, then the section will reach its nominal moment capacity
immediately after cracking; low volumes of reinforcement As tension failure is the intended failure mode φ = 0.9.
will ensure the section is tension-controlled. Defining the Assuming that j = 0.9
internal moment arms of the nonprestressed and prestressed
steel reinforcement for this condition as jd and jdp, respec-   I  1 yt   9 fc′I
Aps 0.8 f ps d p − fse   + e I   + 0.8dAs f y ≥ (13)
tively, the moment capacity can then be defined as   Ayt  final  A 
transfer 
yt

Mn = Ts(jd) + Tps(jdp) (9)


When the member contains only prestressing steel, then
the minimum requirement for bonded prestressed reinforce-
If the extreme tension fiber stresses caused by the prestress
ment is simplified as
force and decompression moment shown in Fig. 5 are
summed, the result is the following equation
9 fc′
Aps,min =
 I  1 yt  y  1 y (14)
0.8  t  f d −f +e t
M dec = Aps fse  
 t  final
Ay  A + e I  (10)  I  final ps p se  A I  transfer
transfer

When the member contains only nonprestressed steel,
This approach results in the self-weight moment being
then the minimum requirement is simplified as
carried by the composite section, which results in a conser-
vative minimum steel requirement. The transfer term should
incorporate the section properties at the time of prestress 11.25 fc′I
As,min ≥ (15)
transfer, and the final term should incorporate the section df y yt

properties of the element after construction is complete.
For composite construction, the transfer term will typi-
To illustrate the implementation of Eq. (14), consider
cally be the noncomposite section properties, and the final
a typical precast bridge with a cast-in-place topping. To
term will be the composite section properties. For noncom-
calculate the minimum reinforcement using the proposed
posite sections, these terms are simply the noncomposite
minimum, the designer would use the moment of inertia,
section properties.
area, and yt of the precast section for the transfer terms,
To ensure sufficient flexural capacity immediately after
as these were the section properties during the prestress
cracking, the resisting moment should be greater than or
transfer. For the final terms, the designer would consider the
equal to the total applied moment. By Leonhardt’s reasoning,
section properties of the completed structure: the composite
the decompression moment is resisted by the already-present
moment of inertia, area, and yt, all considering the cast-in-
prestressing steel; the 1.2 (included to ensure ductility) need
place deck.
only be considered for Mncr. Further, when the safety factor
The proposed provisions are in agreement with the funda-
of 1.2 used by ACI 318-11 and the strength reduction factor
mental approach of providing a moment capacity that is
φ are included, the relationship can be written as
greater than Mcr to ensure a ductile failure—a provision
that has been in the code either explicitly or implicitly
φMn ≥ 1.2Mncr + Mdec (11)
for many years, and has empirically shown itself to be a
successful design practice. The proposed provisions provide
Using Leonhardt’s approach and the strength capacity of
two things: unification of the minimum reinforcement
the prestressing steel beyond the effective prestress to resist
provisions to include both prestressed and nonprestressed
the tensile force in the concrete, the relationship described in
concrete sections (to provide less confusion and a more
Eq. (11) may be rewritten as
uniform level of safety) and revision of the provisions to
eliminate the explicit consideration of the cracking moment

ACI Structural Journal/March-April 2014 435


for prestressed sections (as such, avoiding the problem that
arises in the case of prestressed concrete systems in which
the minimum prestressing cannot be reached.)

PARAMETRIC STUDY: COMPARISON OF


CURRENT AND PROPOSED PROVISIONS
A parametric study was conducted to compare the
minimum area of steel and the subsequent provided capacity
as calculated using the proposed provisions against those
calculated using the ACI 318-11 minimum reinforcement
provisions. A wide range of geometries was chosen for
investigation: standard precast sections (single tees, double
tees, inverted tees, and hollow core slabs), a Florida bulb tee,
a box girder, and an AASHTO girder. Sections were selected
to include geometries commonly used in construction, as
well as both top-heavy and bottom-heavy sections, as the
proposed provisions are dependent on the centroid of the
gross area relative to the centroid of the tension area. Both
nonprestressed and prestressed sections were investigated. Fig. 6—Prestressed noncomposite sections: proposed and
For each cross section, the minimum required area of steel ACI minimums.
was calculated according to ACI 318-11 and the proposed
provisions. The cracking moment of the geometry was bulb tee (FBT78), and a segmental box girder (SB). The
calculated by assuming the concrete tensile strength based SB was considered under negative bending (tension at the
on a modulus of rupture of 7.5√fc′ psi (0.62√fc′ MPa). The top of section). All other sections were considered under
ultimate design strength was calculated using strain compat- positive bending. The nominal moment capacity was calcu-
ibility. The ultimate flexural capacity and the theoretical lated using strain compatibility, and the cracking moment
cracking moment were then compared to determine the was calculated using Eq. (5). Nonprestressed steel was not
margin between cracking and flexural failure. considered in these calculations. Appendix A tabulates data
for all investigated sections.
PRESTRESSED CONCRETE: NONCOMPOSITE A comparison of the ACI minimum and proposed
AND COMPOSITE SECTIONS minimum (from Eq. (14)) reinforcement quantities for
Comparison of the current and proposed provisions is noncomposite sections is shown in Fig. 6. The sections were
awkward because ACI 318-11 minimum reinforcement also evaluated by calculating the ratio of the design strength
provisions (hereafter referred to as ACI minimum) depend φMn to the cracking moment Mcr; this relation is shown in
on a moving target—that is, the cracking moment that Fig. 7. For noncomposite sections, the computed minimum
changes as the amount of bonded prestressing steel changes. area of prestressing steel is within approximately 10% of
On the contrary, the proposed minimum reinforcement the area currently required by the ACI minimum. It can be
provision (hereafter referred to as proposed minimum) does observed that for all sections, the proposed minimum area of
not depend directly on the cracking moment; it depends, steel results in a φMn/Mcr ratio greater than 1, and is close to
instead, on only the material and geometric properties of the 1.2Mcr limit. Notably, comparison of the design moment
the section. Defined as such, the proposed minimum is like strength across a variety of sections indicates a nearly
the AASHTO LRFD minimum reinforcement requirement, constant excess of moment strength (around 17%) beyond
encompassing both types of reinforced sections while basing the cracking limit of each section, ensuring a ductile failure
the minimum steel requirement on the concrete section prop- mode for each section as is intended by the ACI minimum.
erties rather than the cracking moment. Use of the proposed method eliminates the issue that as a
A wide variety of prestressed sections was investigated section increases in depth, the moment capacity increases at
in this parametric study. A table showing the ACI and a faster rate than the cracking moment. As Fig. 8 illustrates, a
proposed minimums for all geometric sections investi- series of varying depth noncomposite hollow core slabs were
gated by this parametric study is included as Appendix A. examined, including a 6 in. (15.2 cm) deep (6HC), an 8 in.
Prestressing strand was assumed to be Grade 270. Prestress (20.3 cm) deep (8HC), a 10 in. (25.4 cm) deep (10HC), and
losses due to elastic shortening and time-dependent losses a 12 in. (30.5 cm) deep (12HC) slab. Appendix A tabulates
were assumed to be 20% after an initial strand stress of these data for all investigated sections. Figure 8 also shows
0.74fpu. The sections in the following discussion were that as the section depth increases, the ratio of the provided
chosen as representative of the range of all standard geome- moment capacity to the cracking moment remains constant.
tries analyzed in the study and include a 12 ft (3.66 m) wide Furthermore, the proposed minimum provides adequate
by 30 in. (76.2 cm) deep double-tee (12DT30), a 28 in. (71.1 reinforcement to ensure ductile failure mode in every case.
cm) wide by 24 in. (61 cm) deep inverted tee (28IT24), a 4 ft A deck was added to each of the noncomposite sections
(1.22 m) wide by 12 in. (30.5 cm) deep hollow core (12HC), evaluated previously (with the exception of the SB) to
an AASHTO IV girder (AIV), a 78 in. (1.98 m) deep Florida investigate the results of the proposed minimum reinforce-

436 ACI Structural Journal/March-April 2014


Fig. 7—Prestressed noncomposite sections: moment
capacity versus cracking moment.
Fig. 9—Prestressed composite sections: proposed and ACI
minimums.

Fig. 8—Prestressed noncomposite hollow core slabs with


proposed minimum.
ment provisions for composite sections. Figure 9 shows the
proposed minimum reinforcement quantity normalized by Fig. 10—Prestressed composite sections: moment capacity
the ACI minimum. The proposed minimum was compared versus cracking moment.
against the ACI minimum calculated for both shored and
unshored construction to demonstrate the full range of typical design volumes of prestressing steel were deter-
expected discrepancies between the two methods. The mined for each of the sections considered in the parametric
sections shown in Fig. 9 were also evaluated by calculating study. Typical design loads were taken from load tables (if
the ratio of the design moment strength φMn to the cracking appropriate), or actual highway designs. Figure 11 shows
moment Mcr using the minimum area of steel prescribed by how the proposed minimum reinforcement compares to the
the proposed method. The results are shown in Fig. 10. volume of reinforcement required by the design. In most
For composite sections, with the exception of the inverted cases, the differences between the current ACI minimum
tee, a 15 to 61% increase in minimum reinforcement over the and the proposed minimum are minor. In general, as can be
ACI minimum was found. The inverted tee is even larger, observed in Fig. 11, the proposed minimum area of steel is
with a 94 to 112% increase. Furthermore, the difference is always less than, and usually much less than, that of a typical
greatest when comparing the proposed minimum to ACI design. Only in the hollow core slab case does the proposed
minimum for an unshored condition. This is, in part, due to minimum approach the typical design value. As Fig. 11
the conservative assumption made in the derivation in which demonstrates, the area of steel is usually controlled by either
the self-weight of the section is assumed to be carried by the stress or strength limits, and the proposed minimum provi-
composite section. To evaluate the impact of this increase, sions will not affect most designs.

ACI Structural Journal/March-April 2014 437


Fig. 11—Proposed area of steel versus typical design. Fig. 12—Nonprestressed sections: proposed and ACI
NONPRESTRESSED CONCRETE: CURRENT minimums.
VERSUS PROPOSED
The current ACI minimum for concrete reinforced with
nonprestressed steel requires a minimum area of steel, as
prescribed by Eq. (2). In general, minimum steel requirements
are usually invoked when sections are larger than required
for strength, such as for architectural or other reasons. To
evaluate the effect the proposed provisions would have in
these instances, the minimum steel requirement was allowed
to control the quantity of steel in the section. The steel quan-
tity was then compared with the current ACI minimum.
Two standard reinforced geometries were analyzed as
nonprestressed concrete, a 10 ft (3.05 m) wide top flange by
48 in. (1.22 m) deep single-tee with an 8 in. (20.32 cm) web
(MS10ST48) and a 28 in. (71.1 cm) wide by 60 in. (1.52 m)
deep inverted tee (MS28IT60), as well as three rectan-
gular sections of varying depth: 12 x 12 in. (30.5 x 30.5
cm), 12 x 36 in. (30.5 x 91.4 cm), and 12 x 72 in. (30.5 x
182.9 cm). All sections were assumed to be under positive
bending (compression in the top of the section). A compar- Fig. 13—Nonprestressed sections: moment capacity versus
ison of ACI and proposed minimums is shown in Fig. 12. cracking moment using proposed provision and ACI 318-11
Appendix B tabulates these data for all investigated sections. (ACI Committee 318 2011) Eq. (10-5).
The sections were also evaluated by calculating the ratio
of the design moment capacity φMn to the cracking moment steel quantity, as inspection of the section’s moment capacity
Mcr; these ratios are shown in Fig. 13. The nominal moment and cracking moment reveals, the proposed method still
capacity was calculated using strain compatibility, and the ensures a ductile failure mode. The rectangular sections and
cracking moment was calculated using Eq. (5). the inverted tee reveal the consistent level of conservatism
Figure 12 highlights the conservative nature of the current achieved by the proposed method with less steel. As calcu-
ACI minimum. In the sections selected for this parametric lated using the proposed provisions, these sections require
study, the ACI minimum for nonprestressed members approximately 15 to 30% less steel than the ACI minimum,
provide between 1.25 and 2.0 safety margins, growing exces- and yet provide a uniform moment capacity near 1.6Mcr.
sively conservative in sections with tensile zones relatively In fact, an inspection of all of the nonprestressed sections
larger than their compressive zones. Parametric studies reveals that the proposed method, unlike ACI, provides a
completed by Seguirant et al. (2010) demonstrated the vari- consistently conservative design with ductile failure mode—
ability of the safety margin provided by the ACI minimum the proposed minimum requirement for each section type,
for nonprestressed sections from slightly under-conservative regardless of the section geometry, provides a capacity equal
to extremely over-conservative (up to 4.43 margin of safety). to approximately 1.6Mcr. Considering this, the proposed
In most cases, the ACI minimum requires more steel than provisions, although requiring less steel in some cases,
the proposed provisions. Despite the general reduction in provide a more reliable estimate of minimum reinforcement
required, and one that is independent of cracking moment.

438 ACI Structural Journal/March-April 2014


SUMMARY Mdnc = unfactored dead load moment acting on monolithic or noncom-
The derivation and parametric studies described in this posite section
Mn = nominal flexural strength at section
paper provide a direct and unified method of determining Mncr = moment applied beyond decompression moment to reach
minimum steel requirements for both nonprestressed rein- rupture stress in extreme tensile fiber; net cracking moment
forced and prestressed sections. For composite prestressed Mr = flexural resistance
Mu = factored moment at section
sections, the moment capacity provided by the area of steel p = ratio of area of tension reinforcement to effective area of concrete
determined with the proposed minimum always exceeds Sc = section modulus with respect to extreme tensile fiber of
1.2Mcr. While the proposed method requires more steel composite section
Snc = section modulus with respect to extreme tensile fiber of mono-
for composite prestressed sections than the ACI 318-11 lithic or noncomposite section
minimum reinforcement provision, the difference is negli- Tps = tensile force in prestressed steel
gible when considering the steel typically provided for Ts = tensile force in nonprestressed steel
yt = distance from centroidal axis of gross section, neglecting rein-
design. For nonprestressed sections, the proposed method forcement, to tension face
results in a consistent moment capacity of 1.6Mcr, regard- φ = strength reduction factor
less of the section shape. Noncomposite prestressed sections
have a consistent moment capacity of 1.17Mcr. The proposed REFERENCES
method has several advantages for a designer; it is indepen- AASHTO, 2007, “AASHTO LRFD Bridge Design Specifications,”
4th edition with 2007 interim revisions, American Association of State
dent of the cracking moment and explicitly calculated. Highway and Transportation Officials, Washington, DC, pp. 5-44 through
5-45.
AUTHOR BIOS ACI-ASCE Committee 323, 2004, “Landmark Series: Tentative Recom-
ACI member Natassia R. Brenkus is a Graduate Research Assistant at the mendations for Prestressed Concrete,” Concrete International, V. 26, No.
University of Florida, Gainesville, FL, where she received her BS in 2006 3, Mar., pp. 95-131.
and her ME in 2012. ACI Committee 318, 1963, “Building Code Requirements for Rein-
forced Concrete (ACI 318-63),” American Concrete Institute, Farmington
H. R. Hamilton, FACI, is the Byron D. Spangler Professor of Civil Engi- Hills, MI, 144 pp.
neering at the University of Florida. He is past Chair of Joint ACI-ASCE ACI Committee 318, 1986, “Building Code Requirements for Rein-
Committee 423, Prestressed Concrete, and a member of ACI Subcommittee forced Concrete (ACI 318-83) (Revised 1986),” American Concrete Insti-
318-6, Precast and Prestressed Concrete (Structural Concrete Building tute, Farmington Hills, MI, 114 pp.
Code). ACI Committee 318, 1992, “Building Code Requirements for Struc-
tural Concrete (ACI 318-92) and Commentary (ACI 318R-92),” American
Concrete Institute, Farmington Hills, MI, 347 pp.
ACKNOWLEDGMENTS ACI Committee 318, 1995, “Building Code Requirements for Struc-
The authors would like to thank C. O’Neill for her contributions to this tural Concrete (ACI 318-95) and Commentary (ACI 318R-95),” American
paper and the Florida Department of Transportation for providing funding Concrete Institute, Farmington Hills, MI, 373 pp.
that supported this work. ACI Committee 318, 2008, “Building Code Requirements for Structural
Concrete (ACI 318-08) and Commentary,” American Concrete Institute,
NOTATION Farmington Hills, MI, 473 pp.
A = area of concrete cross section ACI Committee 318, 2011, “Building Code Requirements for Structural
Act = area of concrete in flexural tension Concrete (ACI 318-11) and Commentary,” American Concrete Institute,
Aps = area of prestressing steel in flexural tension zone Farmington Hills, MI, 503 pp.
A s = area of nonprestressed longitudinal tension reinforcement ACI Committee 501, 1936, “Building Regulations for Reinforced
bw = web width, or diameter of circular section Concrete,” ACI Journal, V. 32, No. 3, Mar., pp. 407-444.
b′ = width of web in I- or T-beam sections Bureau of Public Roads, 1954, “Criteria for Prestressed Concrete
d = depth from compression face to center of nonprestressed tensile Bridges,” U. S. Government Printing Office, Washington, DC, 25 pp.
reinforcement Freyermuth, C. L., and Aalami, B. O., 1997, “Unified Minimum Flex-
dp = depth from compression face to center of prestressed tensile ural Reinforcement Requirements for Reinforced and Prestressed Concrete
reinforcement Members,” ACI Structural Journal, V. 94, No. 4, July-Aug., pp. 409-420.
e = distance between prestressed reinforcement and centroidal axis Ghosh, S. K., 1986, “Exceptions of Precast Prestressed Concrete
fc′ = specified compressive strength of concrete Members to Minimum Reinforcement Requirements,” PCI Journal, V. 31,
fcpe = compressive stress in concrete due to effective prestress forces No. 6, pp. 74-91.
only (after allowance for all prestress losses) at extreme fiber of Khan, A. A.; Cook, W. D.; and Mitchell, D., 1996, “Tensile Strength of
section where tensile stress is caused by externally applied loads Low, Medium, and High-Strength Concretes at Early Ages,” ACI Materials
fpe = compressive stress in concrete due to effective prestress forces Journal, V. 93, No. 5, Sept.-Oct., pp. 487-493.
only (after allowance for all prestress losses) at extreme fiber of Kleymann, M.; Girgis, A.; Tadros, M. K.; and Vranek, C. J., 2006,
section where tensile stress is cause by externally applied loads “Open Forum Problems and Solutions—Minimum Reinforcement in Flex-
fps = stress in prestressing steel at nominal flexural strength ural Members,” PCI Journal, V. 51, No. 5, Sept.-Oct., pp. 146-148.
fr = modulus of rupture of concrete Leonhardt, F., 1964, Prestressed Concrete Design and Construction,
fse = effective stress in prestressing steel (after allowance for all Wilhelm Ernst & Sohn, Germany, 677 pp.
prestress losses) O’Neill, C. M., and Hamilton, H. R., 2009, “Determination of Service
fy = specified yield strength of reinforcement Stresses in Pretensioned Beams,” No. BD 545-78, Florida Department of
I = moment of inertia of section about centroidal axis Transportation.
j = distance between compressive resultant force and tensile resul- Oladapo, I. O., 1968, “Relationship between Moment Capacity at
tant force Flexural Cracking and at Ultimate in Prestressed Concrete Beams,” ACI
Mcr = cracking moment, computed as sum of decompression and net Journal, V. 65, No. 10, Oct., pp. 863-875.
cracking moment, Mdec + Mncr Seguirant, S. J.; Brice, R. B.; and Khaleghi, B., 2010, “Making Sense of
Mdec = applied moment causing the net stress at bottom of section to be Minimum Flexural Reinforcement Requirements for Reinforced Concrete
zero Members,” PCI Journal, V. 55, No. 3, Summer, pp. 64-85.
Walker, S., and Bloem, D. L., 1960, “Effects of Aggregate Size on Prop-
erties of Concrete,” ACI Journal, V. 32, No. 3, Mar., pp. 283-298.

ACI Structural Journal/March-April 2014 439


APPENDIX A
f = 0.9 e, in. Area, in.2 Mn, kip-ft Mcr, kip-ft fMn/Mcr
Table 8X16 2.448 1067.9 629.4 1.53
12DT30 Prop min 10 0.924 411.0 309.5 1.20
ACI min 0.931 414.1 311.0 1.20
Table 18X8 2.750 532.4 506.4 0.95
28IT24 Prop min 2.73 0.481 214.6 165.6 1.17
ACI min 0.520 230.3 171.3 1.21
Table 6X6 0.510 47.2 30.7 1.39
6 in. hollow core Prop min 1.5 0.314 30.1 23.2 1.17
ACI min 0.339 32.3 24.2 1.20
Table 6X6 0.510 70.2 48.6 1.30
8 in. hollow core Prop min 1.5 0.382 53.6 41.1 1.17
ACI min 0.405 56.4 42.3 1.20
Table 4X8 0.612 110.7 77.3 1.29
10 in. hollow core Prop min 1.5 0.467 85.7 65.8 1.17
ACI min 0.499 91.2 68.3 1.20
Table 7X6 0.595 135.1 97.1 1.25
12 in. hollow core Prop min 1.5 0.502 114.2 87.6 1.17
ACI min 0.535 121.1 90.8 1.20
Typical 7.350 3157.8 4162.0 0.68
AASHTO Type IV Prop min 8 1.309 1284.4 990.0 1.17
ACI min 1.405 1373.1 1028.7 1.20
Typical 7.490 11403.3 6704.2 1.53
FBT78 Prop min 8 2.237 3453.8 2611.8 1.19
ACI min 2.290 3533.6 2649.7 1.20
Typical 65.290 209444.4 173166.7 1.09
Box girder Prop min 6.75 45.148 147777.8 115096.7 1.16
ACI min 51.100 166235.6 124574.2 1.20
Notes: 1 in. = 2.54 cm; 1 in.2 = 6.45 cm2; 1 kip-ft = 1.36 kN-m.

APPENDIX B
f = 0.9 Area, in.2 Mn, kip-ft Mcr, kip-ft fMn/Mcr
1.2Mcr 1.310 281.1 210.8 1.20
MS 10ST48 prop min 1.838 393.3 210.8 1.68
ACI Sect 10-5 min 1.358 291.4 210.8 1.24
1.2Mcr 2.895 763.2 570.7 1.20
MS 28IT60 prop min 3.932 1024.4 570.7 1.62
ACI Sect 10-5 min 4.619 1194.4 570.7 1.88
1.2Mcr 0.310 15.2 11.4 1.20
MS 12X12 prop min 0.416 20.7 11.4 1.63
ACI Sect 10-5 min 0.389 18.8 11.4 1.48
1.2Mcr 0.820 138.0 102.4 1.21
MS 12X36 prop min 1.122 187.6 102.4 1.65
ACI Sect 10-5 min 1.300 215.6 102.4 1.89
1.2Mcr 1.580 544.4 409.8 1.20
MS 12X72 prop min 2.188 750.9 409.8 1.65
ACI Sect 10-5 min 2.666 911.7 409.8 2.00
Notes: 1 in. = 2.54 cm; 1 in.2 = 6.45 cm2; 1 kip-ft = 1.36 kN-m.

440 ACI Structural Journal/March-April 2014


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 111-S39

Lateral Strain Model for Concrete under Compression


by Ali Khajeh Samani and Mario M. Attard
The relationship between the lateral and axial strain is important Imran and Pantazopoulou,7 Jamet et al.,8 Lee and Willam,9
when predicting the confinement stresses within reinforced Lu and Hsu,10 Newman,11 and Smith et al.12 The major focus
concrete or fiber-reinforced polymer confined columns. Difficul- of this paper is to establish an analytical relationship between
ties in measuring reliable lateral strains in triaxial compressive lateral and axial strain for concrete under uniaxial compres-
experiments mean that there is a scarcity of lateral strain experi-
sion or compression under confinement. Such a model is
mental results. Two recent lateral strain models will be compared
important when determining the confinement and ductility
with available experimental results. Discussed in this paper is the
transition point in the lateral and axial strain relationship at which of reinforced and FRP confined concrete columns. Further-
the volumetric strain changes sign, and how this transition point is more, specimen size effects associated with the formation of
related to the peak stress. A lateral strain-versus-axial strain model a shear band and tensile cracking during softening are also
is proposed based on the supposition that the concrete behaves incorporated into the model.
linear elastically in the early stages of loading. Once microcracks
form, nonlinear hardening occurs up to the peak stress. After the LATERAL CONCRETE STRAIN UNDER
peak stress, the inelastic lateral strain varies linearly with the COMPRESSION
inelastic axial strain. The lateral-to-axial inelastic strain ratio The lateral strain behavior of concrete under compression
is shown to be a function of the lateral confinement level and the is a key aspect of concrete behavior. Studies on the lateral
failure mechanism. Moreover, the shear band and tensile cracking-
strain behavior under axial compression are conducted under
induced size effect is also discussed from the lateral strain versus
various loading regimes. The loading conditions referenced
axial strain perspective.
herein are mainly uniaxial and triaxial tests conducted on
Keywords: confined concrete; fracture energy; lateral strain; size effect; cylindrical specimens. Triaxial testing of concrete is usually
volumetric strain. conducted in Hoek cells or triaxial cells. In a Hoek cell, the
loading path begins with a sitting or locking axial load. After
INTRODUCTION securing the specimen, the lateral confinement is enforced
There are many design formulas for predicting the up to the desired confinement level. Once the confinement
behavior of reinforced concrete column behavior. Many of reaches the desired value, the axial load is then gradually
these formulas depend on estimating the lateral confinement. applied while the confining stress is maintained. Because the
The lateral ties providing confinement are usually assumed lateral confining stress is applied using liquid pressure and the
to be at yield when a column reaches its peak load-carrying axial strain is externally applied, it is very difficult to enforce a
capacity. For reinforced high-strength concrete columns, hydrostatic load state without the use of complicated comput-
the confinement ties are generally not at yield at the peak erized devices. The triaxial cell used on concrete specimens is
column load-carrying capacity. Therefore, this implies that similar to those used for triaxial testing of soils.
formulas which rely on the confining ties to be at yield are It is important to note the limitations of the various tech-
not applicable. To overcome this problem, the lateral expan- niques used to extract or estimate the lateral strain. Lateral
sion of the confined column core can be related to the stress strains can be measured by utilizing strain gauges on the
in the confining ties, which can then be used to predict the specimens or measuring the deformation of the circumfer-
core confinement. The reinforcing ties are assumed to be ence via a lateral ring. Strain gauges can show some inaccu-
fully bonded to the concrete, with no sliding or slippage racy under high confining pressures inside the test cell. More-
occurring during the loading of the column. The lateral over, the lateral strain can be drastically different at different
deformation of the concrete and the confining reinforcement locations on the specimen. The change in the volume of oil
cage or fiber-reinforced polymer (FRP) wrap is then taken used to apply the confinement can also be used to measure
as equal and related to axial strain in the column core. The the volumetric strain, which can then be used to estimate
lateral strain relationship has been used in the prediction of an average lateral strain. For an intact specimen, the lateral
the confinement level for columns confined by reinforce- strain obtained from the volumetric changes in the specimen
ment cages by Cusson and Paultre1 and Ahmad and Shah,2 correlates well with the lateral strain measured from strain
and FRP wraps by Talaat and Mosalam3 and Lokuge et al.4 gauges and linear variable displacement transducers (LVDTs)
placed at several locations on the surface of the specimen. If
RESEARCH SIGNIFICANCE
To predict the lateral expansion of axially loaded concrete,
ACI Structural Journal, V. 111, No. 2, March-April 2014.
a lateral strain-versus-axial strain relationship is needed. MS No. S-2012-177.R3, doi:10.14359.51686532, was received March 27, 2013, and
reviewed under Institute publication policies. Copyright © 2014, American Concrete
Baseline experimental data on measured lateral strain from Institute. All rights reserved, including the making of copies unless permission is
uniaxial and triaxial loaded concrete are limited to a few obtained from the copyright proprietors. Pertinent discussion including author’s
closure, if any, will be published ten months from this journal’s date if the discussion
research publications such as Candappa et al.,5 Hurlbut,6 is received within four months of the paper’s print publication.

ACI Structural Journal/March-April 2014 441


Fig. 1—Typical: (a) lateral strain; and (b) volumetric strain diagrams for concrete subjected to increasing levels of confine-
ment. (Note: 1 MPa = 145 psi; 1 mm = 0.0394 in.)
localization is present, however, the lateral strains obtained equal to 0.5ε0. The strain ε0 corresponding to the peak
from the changes in oil volume and the strains obtained from stress is given by Eq. (3), and fc′ is the uniaxial compressive
strain measurement devices are drastically different. strength of concrete in MPa.
Lateral strain results are presented as either a set of axial
stress-versus-axial strain and axial stress-versus-lateral   f 
e 0 = e c × 1 + (17 − 0.06 fc′)  r   (fc′ in MPa; 1 MP
Pa = 145 psi) (3)
strain or volumetric strain-versus-axial strain diagrams   fc′ 
(Fig. 1). If volumetric strains are used to obtain the lateral
strain, the obtained lateral strain is an average strain for the In the previous equation, fr is the lateral confinement
whole specimen. Consider, for example, the situation where a on the specimen. The strain at the peak stress for uniaxial
specimen under confinement and axial compression is loaded compression, εc, is taken as 0.002 for all concrete grades.
past the peak and reaches a residual axial load level after In the model proposed by Binici,13 the secant ratio νs
softening. If a localized shear band has formed, the applied is used to correlate the lateral deformation with the axial
axial displacements from the loading machine only impose deformation. The concrete behavior in the lateral direction is
a sliding displacement on the specimen. This displacement divided into three regions: the elastic region, inelastic hard-
does not represent any change of specimen volume, but corre- ening region, and softening region. The model is defined by
sponds to sliding across the shear band. In this case, the actual
change in volume is zero, but the plastic strain or permanent ε′ = νsε (4)
displacement across the whole specimen is not zero. If LVDTs
are used across the fractured zone, a displacement will be in which, ε′ is the lateral strain, ε is the axial strain, and νs is
measured, and a plastic or irrecoverable strain estimated. the secant ratio, defined by

Analytical models by Lokuge et al.4 and Binici13 ν s = νe e ≤ ee


Lateral strain versus axial strain models developed by
Lokuge et al.4 and Binici13 are presented as follows, and   e − ee  2  (5)
νs = νl − ( νl − νe ) exp  −    e > ee
are later compared with some experimental results. These   ∆  
two models are analytical models for the total lateral strain,
and are functions of the axial strain. The model proposed by
Lokuge et al.4 is defined by In the previous equation, νe is the Poisson’s ratio of the
concrete, which varies between 0.15 to 0.2; εe, νl, and Δ are
 a e  defined by
vi   e ≤ e1
e ′   e0   f 
2
= (1) f f 
e 0′   e  a fc′  0.1 1 + 9.9 r − 0.9  r  + r 
e > e  fc′  fc′ fc′
  e  1
ee =
 0 4750 fc′
1
where νl = ν p + 4
  fr   (6)
a = 0.0177fc′ + 1.2818; vi = 8 × 10 (fc′) + 0.0002fc′ + 0.138 (2)
a –6 2   f ′ + 0.85
 c 
(fc′ in MPa, 1 MPa = 145 psi) e0 − ee
∆=
 νl − ν p 
and ε is the axial strain, ε′ is the lateral strain, ε1 can be − ln 
obtained by equating the two right hand side expressions in  νl − ν0 

Eq. (1), and ε0′ is the lateral strain at the peak axial stress

442 ACI Structural Journal/March-April 2014


Fig. 2—Comparison of Jamet et al.8 test results with: (a) Binici13; and (b) Lokuge et al.4 models. (Note: 1 MPa = 145 psi; 1
mm = 0.0394 in.)
Herein, νp is the ratio of lateral strain to the axial strain at was used to impose the desired confinement pressure on the
the peak stress and assumed to be 0.5, νl is the largest secant triaxial test specimens. The confining pressures ranged from
ratio, and ε0 is the axial strain at the peak stress defined by 0.69 to 13.79 MPa (0.1 to 2 ksi). The axial displacements
were transmitted to the specimens via steel rams without
(
e 0 = 5 × −0.067 fc′ 2 + 29.9 fc′ + 1053 ) any friction-reduction measures. Hurlbut6 observed that the
unconfined compression test exhibited the largest dilatant
(7) behavior in the post-peak regime, with the radial displace-
 f f   ment three times greater than the applied axial displace-
×   1 + 9.9 r + r  − 0.8 × 10 −6 ments. The results showed a large volumetric expansion in
 fc′ fc′ 
the post-peak region. In agreement with the observations by
Imran and Pantazopoulou,7 the results of Hurlbut6 showed
Comparing existing models with experimental that as the level of confinement was increased, the dilatancy
results for confined concrete rapidly decreased while exhibiting a transition from large
In this section, the two analytical models described previ- volumetric expansions under low confinement to an elastic
ously are compared with available experimental results. In volumetric compaction for high confinement. The results of
the following comparisons, the elastic moduli and the value Hurlbut6 and Willam et al.14 are compared with the models
of the axial strain at the peak stress used in the analytical of both Binici13 and Lokuge et al.4 in Fig. 3.
models is adopted from the experimental results. The first set Smith et al.12 and Smith15 tested 51 cylindrical samples
of results are from Jamet et al.,8 who tested microconcrete with dimensions of 54 x 108 mm (2.125 x 4.25 in.), which
in triaxial and uniaxial compression. The results of Jamet were tested either in an unconfined or confined state using
et al.8 for lateral strain were extracted from measurements a Hoek cell. The compressive strength was either 34.5 or
of the volumetric strain rather than from LVDTs, strain 44.2 MPa (5.0 or 6.4 ksi). No friction-reducing layer was used
gauges, or both. The cylindrical specimens had a diameter between the steel rams and the specimen ends. The confine-
of 110 mm (4.3 in.) and height of 220 mm (8.7 in.). The ment applied to the specimens varied with a maximum of
uniaxial compression strength was approximately 26 MPa 34.5 MPa (5.0 ksi). The test results on 34.5 MPa (5.0 ksi)
(3.8 ksi). The applied confinement levels were 0, 3, 10, 25, concrete specimens are compared with Binici13 and Lokuge
50, and 100 MPa (0, 0.4, 1.5, 3.6, 7.3, and 14.5 ksi). Jamet et al.4 models in Fig. 4.
et al.8 presented results using true strain rather than engi- The analytical models proposed by Binici13 and Lokuge et
neering or simple strain values. It can be shown that for al.4 give a reasonable fit to the trend of the results in Fig. 2
the magnitudes of strains measured, the true strain and the to 4, but not a good match to the lateral strain quantities.
corresponding engineering strain are virtually indistinguish- Moreover, one can see that neither of the models predicts the
able. All results presented herein are based on engineering initial compaction in the lateral direction under high confine-
strains rather than true strains. The lateral strain results are ment before the specimen starts to expand.
compared with the models proposed by Binici13 and Lokuge The test results of Lee and Willam,9 who tested cylin-
et al.4 in Fig. 2. drical specimens with three different heights in uniaxial
Hurlbut6 and Willam et al.14 tested normal strength compression, are also compared in Fig. 5. All the speci-
concrete in direct tensile and uniaxial compression, as well mens had a diameter of 76.2 mm (3 in.). The heights were
as triaxial compression. The ultimate uniaxial compressive 137.2, 91.44, or 45.72 mm (5.4, 3.6, and 1.8 in.). The peak
strength was 22 MPa (3.2 ksi). The size of the cylindrical concrete strength was 30.7 MPa (4.5 ksi) for the specimens
specimens in the uniaxial compression tests had a diameter with heights of 137.2 and 91.44 mm (5.4 and 3.6 in.), and
of 76 mm (3 in.) and a height of 152 mm (6 in.), while in the 32 MPa (4.6 ksi) for the specimen with a height of 45.72 mm
triaxial tests, the specimens were 54 mm (2.125 in.) in diam- (1.8 in.). The results obtained by Lee and Willam9 are shown
eter by 108 mm (4.25 in.) in height. A modified Hoek cell in Fig. 5, and demonstrate the effects of different specimen

ACI Structural Journal/March-April 2014 443


Fig. 3—Comparison of Willam et al.14 test results with: (a) Binici13; and (b) Lokuge et al.4 models. (Note: 1 MPa = 145 psi;
1 mm = 0.0394 in.)

Fig. 4—Comparison of Smith15 test results with: (a) Binici13; and (b) Lokuge et al.4 models. (Note: 1 MPa = 145 psi; 1 mm =
0.0394 in.)
levels and concrete strengths, and to incorporate size effect
issues.

PROPOSED LATERAL STRAIN MODEL

Lateral strain at peak axial stress


Kotsovos and Newman16 proposed that the point where
the volume expansion begins is at the onset of unstable frac-
ture propagation (OUFP), and that this should be consid-
ered as the true strength indicator of concrete. The load at
OUFP is usually very close to the peak load (approximately
90%), and it is observed that deformations increase signifi-
cantly between the OUFP and the peak. Based on experi-
mental results, Nielsen17 suggested that the axial strain at
Fig. 5—Comparison of Lee and Willam9 test results compared
the peak stress is 2.2 times greater than the absolute value of
with Binici13 model and Lokuge et al.4 model. (Note: 1 mm =
the lateral strain at peak stress. Imran18 studied the effect of
0.0394 in.)
water-cement ratio, concrete strength, and confinement level
height on the measured lateral strains. Figure 5 also shows on the lateral strain of concrete. Figure 6 shows the relation-
the comparison with the models of Binici13 and Lokuge et ship between the axial compressive strain at the peak stress
al.4 The lateral strain size effect was not considered in the ε0, and the axial compressive strain at the instant of zero
model of Lokuge et al.4 Binici13 did consider specimen size volumetric strain. A similar figure was originally presented
effect on the axial strain-versus-axial stress relationship, but by Imran.18 Herein, results from the experimental studies
not on the lateral strain, as the model of Binici13 considers of Candappa et al.,5 Jamet et al.,8 Newman,11 and Smith15
the lateral strain as a direct function of the axial strain. A new have been augmented to Imran’s results. The diagonal line in
model for predicting the lateral strain is developed in the Fig. 6 would correspond to a situation where the volumetric
following section, which attempts to give a better predictor strain is zero at the instance of the peak load. Imran and
of the lateral strain for a broad spectrum of confinement Pantazopoulou7 concluded that the axial compressive strain

444 ACI Structural Journal/March-April 2014


Fig. 6—Axial strain at zero volumetric change versus axial Fig. 7—Plastic lateral strain minus plastic lateral strain
strain at peak stress. (Note: 1 mm = 0.0394 in.) at the peak stress versus the plastic axial strain minus the
plastic axial strain at the peak stress for a 40 MPa concrete
at the instant of zero volumetric strain is approximately specimens tested by Candappa et al.5 (Note: 1 MPa =
equal to the axial compressive strain at which the concrete 145 psi; 1 mm = 0.0394 in.)
reaches its peak compressive axial stress. The volumetric
strain εv at the peak axial stress can be written as in which εp is the plastic axial strain, and εp′ is the plastic
lateral strain. The quantities εp(at peak) and εp′(at peak) are the
ev = e 0 + 2 e 0′ = 0 (8) axial and lateral plastic strain at the peak stress level, respec-
e = e0
tively. If it is assumed that the elastic part of the strains can
be calculated using Hooke’s law, then Eq. (10) becomes
where ε0′ is the lateral strain at the peak axial stress. Hence,
the ratio of the lateral strain ε0′ to the axial strain at the peak  f f    f f 
stress μ0 can be taken as e ′ −  r − µ e .  −  e 0′ −  r − µ e . 0  
 Ec Ec    Ec Ec  
(11)
μ0 = ε0′/ε0 = –0.5 (9)   f f    f f  
= b e −  − µe . r  −  e0 −  0 − µe . r   
  Ec Ec    Ec Ec   
This assumption will be used in this study.

Post-peak plastic lateral strain Using Eq. (9), the previous equation becomes
Vonk19 observed that in the first stage of loading, the
average lateral deformation versus average axial defor-  f    f 
mation is governed by Poisson’s ratio, which results in a e ′ +  µ e ⋅  − µ 0 e 0 +  µ e ⋅ 0  
decrease of the volume of the specimen. In the second stage,  Ec    Ec  
(12)
macrocracks are formed, causing dilatant behavior. After a   f    f  
gradual change, a more-or-less constant dilatant behavior = b e −   −  e0 −  0   
was observed, which indicated that the process of macroc-   Ec    Ec   
rack formation had stopped, and the final crack pattern had
been developed. The behavior of concrete after the peak In the previous equation, f is the stress level at a strain of ε;
stress can be quantified by looking at the plastic or inelastic f0 is the peak stress, and Ec is the elastic modulus of concrete.
deformations after the peak. Figure 7 plots the plastic lateral The axial strain versus axial stress model for confined
strain minus the plastic lateral strain at the peak stress against concrete proposed by Samani and Attard20 will be used for
the plastic axial strain minus the plastic axial strain at the the relationship between the axial strain and axial stress.
peak stress based on the test results of Candappa et al.1 The Experimental estimates for the parameter β as a function
results show an almost linear relationship. For relatively low of the confinement ratio, obtained from the tests of Candappa
levels of confinement associated with the results in Fig. 7, et al.,5 Hurlbut,6 Jamet et al.,8 Lu and Hsu,10 Newman,11 and
a shear band has probably formed with the deformations in Smith15 are shown in Fig. 8. The experimental data used in
the axial and lateral directions concentrated on the deforma- Fig. 8 are for specimens with a height-diameter or height-
tions within the shear band. The slope of the linear trends in width ratio of h/D or h/w = 2. The experimental results in
Fig. 7 are associated with differing levels of confinement. Fig. 8 show a large scatter, particularly for the uniaxial
As the confinement level increases, the lateral deformation compression state. The β parameter varies from approxi-
rate decreases. An equation for this approximate linear trend mately –2.5 (for h/D or w/h = 2) at the uniaxial state, to an
is therefore upper limit of –0.5 for high confinement.
To gain an understanding of the numerical range for the
εp′ – εp′(at peak) = β[εp – εp(at peak)] (10) parameter β, consider a simple Mohr Coulomb nonassoci-

ACI Structural Journal/March-April 2014 445


residual level under confinement, the dilatancy angle would
be zero if a shear band forms and there is purely friction type
sliding across the shear band. When the dilatancy angle is
zero, Eq. (15) would give

e ′p
b≈ = −1 (16)
e p

Referring to Fig. 8, the β parameter is approximately –1


when the confinement ratio is approximately 10 to 20%.
In the study by Samani and Attard,20 it was observed that
the post-peak compressive fracture energy is not constant,
but varies with the level of confinement. The compressive
Fig. 8—Variation of parameter β with confinement. (Note: fracture energy increases with increasing confinement and
1 MPa = 145 psi.) reaches a limit at a confinement ratio of approximately 10
to 20%. The compressive fracture energy then decreases
ated plasticity model. The model has an initial failure angle for increasing confinement until the compressive fracture
φf, and a plastic strain dilatancy angle ψ. Assuming there is energy becomes zero after the transition from brittle to hard-
a shear failure plane with an inclination angle θ measured ening failure. It was suggested by Samani and Attard20 that
from the horizontal, the inclination of the failure shear plane for uniaxial compressive and triaxial compression with low
can be related to the Mohr Coulomb failure surface by levels of confinement, tensile cracking exists along with the
formation of a shear band during failure (Fig. 9). In Fig. 9,
π ff hd is the damage zone height taken to be between 2 to 2.5
θ= + (13)
4 2 times the width or diameter of the specimen. With increasing
levels of confinement, eventually tensile splitting is nulli-
The plastic strain rate acting on the shear failure plane can fied, and the compressive fracture energy increases above
be decomposed vectorially into a plastic normal strain rate, the uniaxial level. As the confinement is further increased,
and a plastic shear strain rate. Incorporating a nonassociated the mode of failure changes, and is dominated by barreling
flow rule, the relationship between the normal and shear dispersed cracking and the compressive fracture energy
plastic strain rates is decreases to zero. Applying these observations to the lateral
strain phenomenon, the presence of tensile splitting coupled
e np with a shear band failure under uniaxial and low confine-
tan ( ψ ) = (14) ment could explain why the measured β parameter is much
e sp
greater than predicted using Eq. (15) based on a simple Mohr
Coulomb failure criterion and the assumption of a single
The relationship between the plastic axial strain rate and shear band. Tensile splitting increases the magnitude of the
plastic lateral strain rate is then defined by relative lateral dilation. Vonk19 observed that “the formation
of a larger number of splitting cracks causes larger lateral
e ′p tan ( ψ − θ) deformations.” The plotted data for the parameter β in Fig. 8
= (15) is categorized in Fig. 10 to reflect the failure mode. When
e p tan (θ)
the confinement ratio is less than 10 to 20%, the failure and
fracture is due to tensile splitting and shear failure, with the
For concrete, the value of the dilatancy angle is approx- associated β less than –1. As the confinement increases,
imately 13 degrees.21 The Mohr Coulomb failure surface tensile splitting is nullified, and failure is predominately a
angle (before softening) is usually approximately 30 degrees, shear band failure with an associated β of between –1 to
which results in a shear plane failure angle of approximately –0.5. Increasing the lateral confinement beyond the brittle
60 degrees. Using Eq. (15), the ratio of the plastic lateral continuous hardening transition point (confinement ratio of
strain rate to the plastic axial strain rate is then roughly approximately 30 to 50%), the concrete behavior changes
–0.58. The ratio of the plastic lateral strain rate to the plastic from shear failure to ductile behavior. When ductile, the β
axial strain rate approximates the definition of the parameter parameter is –0.5.
β defined by Eq. (10). The experimentally measured expan- Figure 11 plots experimental estimates for the parameter β
sion rate, as reflected by the parameter β shown in Fig. 8, for for uniaxial compression as a function of the aspect ratio D/h
the uniaxial, low confinement case, or both, is much greater or w/h, where D is the diameter for a cylindrical specimen,
than predicted by assuming a shear plane failure and a Mohr w is the width for a prism, and h is the specimen height.
Coulomb failure surface. The estimate based on Eq. (15) The experimental results are from the work of Lee and
assumed the plastic strain dilatancy angle and the friction Willam,9 Van Mier,22 Hurlbut,6 Lu and Hsu,10 Newman,11
angle to be constant during softening, while it is known and Smith.15 As mentioned previsously, Lee and Willam9
that these parameters change. When the load reduces to a tested cylindrical specimens in uniaxial compression with

446 ACI Structural Journal/March-April 2014


Fig. 10—Concrete behavior based on β equation.

Fig. 9—Shear band failure plane and tensile splitting under


uniaxial compression.
three different heights of 137.2, 91.44, and 45.72 mm (5.4,
3.6, and 1.8 in.), and a diameter of 76.2 mm (3 in.). Results
by Van Mier22 are also shown. Van Mier22 tested prisms with
a cross section of 100 x 100 mm (4 x 4 in.) and varying
heights of 50, 100, and 200 mm (2, 4, and 8 in.). Both Lee
and Willam9 and Van Mier22 measured the axial and lateral
strain for specimens with aspect ratios h/D or h/w less than
or equal to 2. The results show a large scatter, which could
be attributed to the increasing dominance of tensile cracking
for specimens with aspect ratios less than 2. Fig. 11—Variation of parameter β for uniaxial compression
An equation for the parameter β defined by Eq. (12) is with varying aspect ratio. (Note: 1 MPa = 145 psi.)
estimated using an analysis of the experimental results by the plastic/inelastic axial strain. A preliminary and primi-
Candappa et al.,5 Hurlbut,6 Jamet et al.,8 Lee and Willam,9 tive version of the model presented herein was presented in
Lu and Hsu,10 Newman,11 Smith et al.,12 Smith,15 and Van Samani and Attard.23 The model described herein is a refined
Mier22 (Fig. 8 and 11). The parameter β is taken as a func- version based on comprehensive research on reinforced
tion of the specimen aspect ratio, concrete strength, and concrete columns, and also includes size effect issues. The
the confinement ratio, but independent of axial strain. The equations for the lateral strains in the three phases are
proposed expression is written as
 f µ f
c e ′ = pr − e e ≤ e pr
b= b
− 0.5;  E c Ec
f  
a ×  r’  + 1   f pr µ e f 
 fc  e ′ =  −  + A1 ( e ) e pr < e ≤ e 0 (18)
  Ec Ec 
−0.015 fc′ −0.02 fc′
a = 65e ; b = 1.5 − e ( fc′ in MPa, 1 MPa = 145 psi) (17) 
e ′ = −µ f + b  e − f −  e − f0   + µ e + µ f0 e > e0
 D  Ec  Ec  
e 0 0 0 e
c = −4   h ≤ hd ; c = −2 h > hd  Ec  Ec
 h

in which ε is the axial strain and ε′ is the lateral strain; μe


Lateral strain model is the elastic Poisson’s ratio that normally varies between
A model for the lateral strain versus axial strain is 0.15 and 0.25; fpr and εpr are the axial stress and strain at
proposed herein based on the assumption that the concrete the proportional limit in concrete (this is usually taken at
behavior is firstly linear elastic. Once the microcracks form, 45% of the peak stress f0); and Ec is the elastic modulus.
the behavior becomes nonlinear hardening up to the peak ε0 is the axial strain at the peak stress; μ0 is the ratio of the
stress. After the peak stress, volumetric expansion starts lateral strain to the axial strain at the peak stress; β is defined
with plastic/inelastic lateral strain varying linearly with by Eq. (17); and A1(ε) is a function describing the dilation

ACI Structural Journal/March-April 2014 447


during the nonlinear hardening phase up to the peak stress. Gft = 0.00097fc′ + 0.0418 N/mm (N/mm = 5.71 lpf/in.) (23)
The Appendix contains the detailed derivation for A1(ε),
which is determined to ensure continuity between the three Samani and Attard20 back-calculated an expression for the
phases in Eq. (18). The lateral strain axial strain model is localized inelastic displacement using the results of Attard
then given by and Setunge26 for a reference cylinder height of hr = 200 mm
(8 in.), that is
 f pr µ e f
e ′ = − e ≤ e pr


Ec Ec
 e − e pr 
2
  e − e pr   w pc = ( e − e 0 ) hr +
( f0 − f ) h − e d hr (24)
 exp  a  r
  e − e  
  e o − e pr  
Ec

e ′ = 
 f pr µ e . f    f pr µ e f0   o pr 
−  + µ 0 e 0 −  −  e pr < e ≤ e 0
  Ec Ec    Ec Ec   exp(a )



 
 e ′ = − µ e f + b e − f − e − f0 + µ e + µ e f0 
 e > e0 Substituting Eq. (21) into (19) for the case of ε > ε0, one
 Ec  Ec  
0 0 0
Ec  Ec
 obtains


(19) f  w pc + e d hd  f0
e ′ = −µ e .
Ec
+ b. 
 h  + µ 0 e 0 + µ e ⋅ E h > hd
The parameter α is given by c
(25)
f  w pc  f
e ′ = −µ e . + b.  + ed  + µ0 e0 + µe ⋅ 0 h ≤ hd
  h  Ec  h  Ec
 b  h  
a= r 

 e ′po  o
(
e − e pr − 2 (20) ) Using Eq. (24), Eq. (25) can be rewritten to give the lateral
  strain for the post-peak region as:
 

f  h ( f − f ) hr  h h 
where hr is the reference cylinder height. To adjust the stress e ′ = −µ e + b  (e − e0 ) r + 0 − ed  r − d  
Ec  h Ec h  h h 
strain model to incorporate size effects, Samani and Attard20
divided the total axial strain ε into its inelastic and elastic f0
components, such that +µ0 e0 + µe h > hd and e > e 0
Ec
(26)
f  hr ( f0 − f ) hr  hr  
 f   w pc + e d hd  f e ′ = −µ e + b  (e − e0 ) + − e d  − 1 
e =  e0 − 0  +   + E h > hd Ec  h Ec h h 
 Ec   h c

(21) f0
 f   w pc  f +µ0 e0 + µe h ≤ hd and e > e 0
e =  e0 − 0  +  + ed  + h ≤ hd Ec
 Ec   h  Ec

Herein, wpc is the localized inelastic axial displacement Figures 12 to 15 show a comparison of the proposed model
due to shear band fracture; hd is the damage zone height; with test results presented in Candappa et al.,5 Hurlbut,6 Imran
and εd is the additional inelastic axial strain in the damaged and Pantazopoulou,7 Jamet et al.,8 and Smith.15 Generally,
zone associated with longitudinal tensile cracking given in the proposed model makes very good predictions giving the
Samani and Attard20 by correct trends, and, in most cases, reasonable estimates of
0.8
the lateral strain. These comparisons demonstrate the capa-
2 kG ft  f0 − f  bility of the proposed model in modelling a wide range of
ed = fresidual ≤ f ≤ f0 (22)
r (1 + k ) f0  f0 − fresidual  compressive strengths and confining pressures. Figures 16
and 17 show comparisons of the new analytical model with
uniaxial compression tests involving specimens of different
where Gft is the tensile fracture energy; r is a parameter with dimensional aspect ratios. The tests of Lee and Willam,9
the dimension of length proportional to the average distance who tested cylindrical specimens in uniaxial condition with
between successive longitudinal cracks; k is a material different heights of 137.2, 91.44, and 45.72 mm (5.4, 3.6,
constant; and fresidual is the residual axial stress level. The and 1.8 in.) and a diameter of 76.2 mm (3 in.), have already
value of r was estimated by Markeset and Hillerborg24 to be been mentioned. Figure 16 shows the comparison with Lee
approximately 1.25 mm (0.05 in.) for a maximum aggregate and Willam9 results. Although the match is only fair, the new
size of 16 mm (0.6 in.), with r increasing with increasing model has the correct trend that shows larger lateral strains
maximum aggregate size. The value of k suggested by for the specimens with the smallest height. Van Mier,27 a
Markeset and Hillerborg24 was approximately 3 for normal pioneer in the work of strain softening and size effect issues,
density concrete and 1 for lightweight aggregate concrete. also presented experimental lateral strain versus axial strain
An expression for the tensile fracture energy as a function results for prisms with a square cross section of 100 x
of the uniaxial compressive proposed by Van Mier25 can 100 mm (4 x 4 in.) and heights of 50, 100, and 200 mm (2, 4,
be used

448 ACI Structural Journal/March-April 2014


Fig. 12—Comparison of lateral strain versus axial strain for
Candappa et al.5 test results (fc′ = 100 MPa) with proposed Fig. 15—Comparison of lateral strain versus axial strain for
model. (Note: 1 MPa = 145 psi; 1 mm = 0.0394 in.) Smith15 test results (fc′ = 34.5 MPa) with proposed model.
(Note: 1 MPa = 145 psi; 1 mm = 0.0394 in.)

Fig. 13—Comparison of lateral strain-versus-axial strain


test results of Willam et al.14 with proposed model. (Note:
1 MPa = 145 psi; 1 mm = 0.0394 in.)

Fig. 16—Comparison of lateral strain versus axial strain for


Fig. 14—Comparison of lateral strain versus axial strain Lee and Willam9 test results with proposed model. (Note:
for Jamet et al.8 test results with proposed model. (Note: 1 MPa = 145 psi; 1 mm = 0.0394 in.)
1 MPa = 145 psi; 1 mm = 0.0394 in.)
mens having aspect ratios less than 2. The proposed model
and 8 in.) under uniaxial compression. Comparisons with the gives quantitatively good results.
results of Van Mier27 are shown in Fig. 17. Van Mier27 tested
several specimens, and presented a range of results for each CONCLUSIONS
specimen height. There is a wide scatter in the test results Two analytical models for the lateral strain versus axial
indicative of the dominance of tensile splitting with speci- strain proposed by Binici13 and Lokuge et al.4 were reviewed
and compared with test results. The correlation between

ACI Structural Journal/March-April 2014 449


NOTATION
A1(ε) = function in proposed model
a = parameter in proposed model
b = parameter in proposed model
c = parameter in proposed model
D = diameter of specimen
Ec = elastic modulus of concrete
f = stress level at a strain of ε
f0 = peak stress
fc′ = uniaxial compressive strength of concrete, in MPa
fpr = axial stress and strain at proportional limit in concrete
fr = lateral confinement on specimen, in MPa
fresidual = residual axial stress level
Gft = tensile fracture energy
h = height of specimen
hd = damage zone height
hr = reference cylinder height
k = material constant
r = parameter with dimension of length proportional to average
Fig. 17—Comparison of lateral strain versus axial strain for distance between successive longitudinal cracks
Van Mier24 test results with proposed model. (Note: 1 MPa = wpc = localized inelastic axial displacement due to shear band fracture
α = parameter in proposed model
145 psi; 1 mm = 0.0394 in.) β = parameter defining the relationship of plastic lateral and axial
strain
the test results and the analytical model predictions was Δ = scaling parameter in Binici13
reasonable, but not for high confinement levels. The initial ε = axial strain
εc = strain at peak stress for uniaxial compression tests
compaction under high confinement was not predicted, and εe = elastic limit strain
the models did not incorporate size effects associated with εe = additional inelastic axial strain in the damaged zone
varying specimen aspect ratios. εp = plastic axial strain
εpr = axial strain at proportional limit in concrete
A new model for the lateral strain-versus-axial strain rela- εp(at peak) = axial plastic strain at peak stress level
tionship has been proposed based on the assumption that ε0 = strain corresponding to peak stress
the concrete behavior could be classified into three phases. ε0′ = lateral strain at peak axial stress equal
ε1 = transition point in Lokuge et al.4
It was assumed that the concrete responds to loads linearly ε′ = lateral strain
elastically up to a proportional limit, and then its response εp′ = plastic lateral strain
changes to a nonlinear hardening up to the peak stress. From ε′p(at peak) = lateral plastic strain at peak stress level
previous studies, the point at which the volumetric strain e np = plastic normal strain rate
becomes zero was taken to be the same as the peak stress e sp = plastic shear strain rate
φf = initial failure angle
point; therefore, the lateral strain at the peak axial stress was μ0 = ratio of lateral strain to axial strain at peak stress
half that of the axial strain at the peak axial stress. In the μe = elastic Poisson’s ratio
post-peak phase, a linear relationship between the plastic νe = Poisson’s ratio of concrete varying between 0.15 to 0.2
νl = largest secant ratio
lateral strain and the plastic axial strain was observed. The νp = ratio of lateral strain to axial strain at the peak stress
total lateral strain was calculated by adding the elastic and νs = secant Poisson’s ratio
estimated plastic components of the lateral axial. The param- θ = shear failure plane
ψ = plastic strain dilatancy angle
eters affecting the relationship between the plastic lateral
and axial strain was shown to be the confinement level and
REFERENCES
the specimen aspect ratio. 1. Cusson, D., and Paultre, P., “Stress-Strain Model for Confined High-
To show the model’s accuracy, the proposed model predic- Strength Concrete,” Journal of Structural Engineering, ASCE, V. 121,
tions were compared with a vast range of results in which 1995, p. 468.
2. Ahmad, S., and Shah, S., “Stress-Strain Curves of Concrete Confined
the concrete strengths varied from low to high strength, and by Spiral Reinforcement,” ACI Journal, V. 79, No. 6, Nov.-Dec. 1982,
differing levels of confinement were applied. The model pp. 484-490.
showed a realistic match with similar trends to experi- 3. Talaat, M., and Mosalam, K., “Computational Modeling of Progres-
sive Collapse in Reinforced Concrete Frame Structures,” PEER Report
mental results. The model also compared reasonably well 2007/10, Pacific Earthquake Engineering Research Center, University of
for uniaxial specimens of different aspect ratios, although California, Berkeley, CA, 2008, 308 pp.
the experimental results showed a large scatter because of 4. Lokuge, W.; Sanjayan, J.; and Setunge, S., “Stress-Strain Model for
Laterally Confined Concrete,” Journal of Materials in Civil Engineering,
the effects of tensile splitting. V. 17, No. 6, 2005, pp. 607-616.
5. Candappa, D.; Sanjayan, J.; and Setunge, S., “Complete Triaxial
AUTHOR BIOS Stress-Strain Curves of High-Strength Concrete,” Journal of Materials in
Ali Khajeh Samani is currently a Post-Graduate Fellow at the School of Civil Engineering, V. 13, No. 3, 2001, pp. 209-215.
Civil and Environmental Engineering, the University of New South Wales, 6. Hurlbut, B., Experimental and Computational Investigation of
Sydney, Australia. His research interests include modeling of confined Strain-Softening in Concrete, University of Colorado, Boulder, CO, 1985,
concrete. 256 pp.
7. Imran, I., and Pantazopoulou, S., “Experimental Study of Plain
Mario M. Attard is an Associate Professor at the School of Civil and Envi- Concrete under Triaxial Stress,” ACI Materials Journal, V. 93, No. 6,
ronmental Engineering, the University of New South Wales. His research Nov.-Dec. 1996, pp. 589-601.
interests include buckling analysis, hyperelastic constitutive modeling, 8. Jamet, P.; Millard, A.; and Nahas, G., “Triaxial Behaviour of
finite strain continuum mechanics, and softening and constitutive modeling Micro-Concrete Complete Stress-Strain Curves for Confining Pressures
of concrete. Ranging from 0 to 100 MPa,” RILEM-CEB International Conference:
Concrete under Multiaxial Conditions, V. 1, May 1984, pp. 133-140.

450 ACI Structural Journal/March-April 2014


9. Lee, Y., and Willam, K., “Mechanical Properties of Concrete in   f pr µ e f0  
Uniaxial Compression,” ACI Materials Journal, V. 94, No. 6, Nov.-Dec. e ′ =  −  + A1 ( e 0 ) 
1997, pp. 457-471.   Ec Ec  
10. Lu, X., and Hsu, C., “Stress-Strain Relations of High-Strength
e = e0 ⇒   ⇒ A1 ( e 0 ) = e ′p0
  f pr µ e f0   f0  f0   
Concrete under Triaxial Compression,” Journal of Materials in Civil Engi- e ′ =  −  + b  e − − e −  + e ′
 Ec  Ec   0 
0 0 p
neering, V. 19, No. 3, 2007, pp. 261-268.   Ec Ec   

11. Newman, J., “Concrete under Complex Stress,” Developments
in Concrete Technology—I, F. Lydon, ed., Applied Science Publishers, (A2)
London, 1979, pp. 151-219.
12. Smith, S.; Willam, K.; Gerstle, K.; and Sture, S., “Concrete Over the The slopes at these transition points also need to be equal,
Top—Or, is there Life After Peak?” ACI Materials Journal, V. 86, No. 5,
Sept.-Oct. 1989, pp. 491-497. therefore
13. Binici, B., “An Analytical Model for Stress-Strain Behavior
of Confined Concrete,” Engineering Structures, V. 27, No. 7, 2005,  de′ µ df 
pp. 1040-1051.  =− e = −µ e 
14. Willam, K.; Hurlbut, B.; and Sture, S., “Experimental and Constitu-  d e Ec d e e = e pr  dA1 ( e )
e = e pr ⇒  → =0
dA1 ( e )
tive Aspects of Concrete Failure,” Finite Element Analysis of Reinforced
Concrete Structures, (Proceedings of seminar sponsored by Japan Society  de ′ = − µe df  d e e = e pr
+
for the Promotion of Science and the U.S. National Science Foundation),  de Ec d e e = e pr d e e = e pr 
1986, pp. 226-254. 
15. Smith, S. H., “On Fundamental Aspects of Concrete Behavior,” MS (A3)
thesis, CEAE Department, University of Colorado Boulder, Boulder, CO,
1987.
16. Kotsovos, M., and Newman, J., “A Mathematical Description of the
Deformational Behavior of Concrete under Complex Loading,” Magazine   0

of Concrete Research, V. 31, No. 107, 1979, pp. 77-90.  de′ µ e df dA1 ( e ) 
17. Nielsen, C., “Triaxial Behavior of High-Strength Concrete and  =− + 
 d e E c d e e = e0 d e e = e0 
Mortar,” ACI Materials Journal, V. 95, No. 2, Mar.-Apr. 1998, pp. 144-151.
 dA ( e ) 
18. Imran, I., “Applications of Non-Associated Plasticity in Modeling = 1

the Mechanical Response of Concrete,” PhD thesis, University of Toronto,  d e e = e0 
Toronto, ON, 1994.   dA1 ( e )
     0

  h
19. Vonk, R., “Softening of Concrete Loaded in Compression,” thesis, e = e0 ⇒  0
⇒ =b
1992.  µ e df  df   d e e = e0 hr
 − + b 1 −  
20. Samani, A., and Attard, M., “A Stress-Strain Model for Uniaxial and Ec d e e = e 0  Ec d e e = e 0 
Confined Concrete under Compression,” Engineering Structures, V. 41,  
 d e ′ 
  h
 =b 
2012, pp. 335-349. =
21. Vermeer, P. A., and De Borst, R., “Non-Associated Plasticity for
 de hr  hr  df hr 
 + 1 −  
Soils, Concrete and Rock,” Technical University of Delft, Delft, Nether-  h  h  E c d e e = e0 
     
lands, 1984.
 0
22. Van Mier, J., “Strain-Softening of Concrete under Multiaxial Loading
Conditions,” Dissertation, 1984, 349 pp. (A4)
23. Samani, A., and Attard, M., “Lateral Behaviour of Concrete,” Inter-
national Conference on Earthquake and Structural Engineering (ICESE
2011), Venice, Italy, 2011, pp. 940-945. A form for the function A1(ε), which allows the continuity
24. Markeset, G., and Hillerborg, A., “Softening of Concrete in Compres- conditions to be satisfied, is
sion—Localization and Size Effects,” Cement and Concrete Research,
V. 25, No. 4, 1995, pp. 702-708. 2
25. Van Mier, J., Fracture Processes of Concrete, CRC Press, Boca  e − e pr    e − e pr  
Raton, FL, 1996, 464 pp.  e − e  exp  a  e − e  
26. Attard, M., and Setunge, S., “Stress-Strain Relationship of Confined  o pr    o pr  
and Unconfined Concrete,” ACI Materials Journal, V. 93, No. 5, Sept.-Oct. A1 ( e ) = e ′p = e ′p0 (A5)
1996, pp. 432-442. exp(a )
27. Van Mier, J., “Multiaxial Strain-Softening of Concrete,” Materials
and Structures, V. 19, No. 3, 1986, pp. 179-190.
with α defined as
APPENDIX A
The proposed lateral strain model continuity and boundary   h 
conditions are detailed here. To define A1(ε) used in Eq. (18),  b  h  
the following continuity conditions need to be satisfied. The a= r 

 e ′p0  o
(
e − e pr − 2 (A6) )
lateral strains at the transition points between the elastic phase  
and the nonlinear hardening phase, and the nonlinear hard-  

ening phase and at the peak stress, need to be equal, hence:

 f µ e ⋅ f pr  
e ′ =  r −  
  Ec Ec  
e = e pr ⇒   → A1 (e pr ) = 0
  f µ e ⋅ f pr  
e ′ =  E − E  + A1 (e pr ) 
r

  c c  
(A1)

ACI Structural Journal/March-April 2014 451


NOTES:

452 ACI Structural Journal/March-April 2014


DISCUSSION
Disc. 110-S31/From the May-June 2013 ACI Structural Journal, p. 404.

Cyclic Loading Test for Beam-Column Connection with Prefabricated Reinforcing Bar Details. Paper by
Tae-Sung Eom, Jin-Aha Song, Hong-Gun Park, Hyoung-Seop Kim, and Chang-Nam Lee
Discussion by Yun Liu and Dun Wang
Lecturer, College of Architecture and Civil Engineering, Zinjiang University, Urumqi, China; ACI member, PhD candidate, Research Institute of Structural Engineering and Disaster
Reduction, School of Civil Engineering, Tongji University, Shanghai, China.

The subject of precast concrete systems has been a 374.2R-13,18 gravity loads should be simulated during
promising research field since the PRESSS (Precast Seismic testing, whether their effects are deemed important
Structural System Program) was initiated by the U.S. and or not, because there is an important aspect of the
Japan in the 1990s.16 Numerous research activities have application of gravity loads on a column, resulting in a
been conducted and many novel structural systems, such as fast rate of lateral force strength degradation. Therefore,
hybrid frame systems, rocking frame systems, and rocking it is best to include the axial load’s effect into the test of
wall systems, have been developed and applied to practical columns to get a more realistic behavior.
projects. The major features of precast concrete systems are • In the “Test Program” section of the paper, it can be
the fast speed of construction, the high quality of precast seen in Table 1 that there are more bottom reinforcing
and prestressed concrete units, durability improvements, bars than top reinforcing bars in the beam, which may
and superior performance during earthquakes.17 To not be the case in practice. What is the consideration
improve the speed of construction and ensure the seismic of such an arrangement by the authors? A second
performance of precast systems, various techniques and question regards the setup of the test. It is indicated
construction methods have been developed, such as the that the ends of the beam cannot move upward and
welded reinforcement grid (WRG) and the SEN Steel downward; instead, it allows the beam end to move
Concrete Construction method, as mentioned in the paper. horizontally. Does such a test setup conform to the
The discussed paper presents a prefabricated reinforcing true situation? What is the consideration for such a
bar connection method for the earthquake design of beam- test setup? Please clarify. Moreover, it is mentioned in
column connections using the techniques of reinforcing the paper that D25 bars used for the beam bottom bars
bar welding, coupler splicing, and headed-bar anchorage. had a relatively small fracture strain—5.36%—which
Test results of four test specimens under cyclic loading are was less than the minimum requirement specified in
given to elaborate on the effectiveness of this construction the Korean Industrial Standard. Strictly speaking, it is
method, which could be an alternative for the application forbidden to use unqualified material in tests and in the
of precast concrete frame systems. Some topics in the paper practical project. However, such steel bars are still in
are interesting and the discussers would like to comment on use for the test specimens, which may be caused by the
them as follows: sequence of the test specimens’ construction—that is,
• Welding stress, cracks, pores, and slags in the weld zone the test specimens were constructed before the material
are complex phenomena for steel bars with welding properties test finished. Although there are requirements
during loading, which have an adverse effect on the for steel bars before entering into the laboratory, it is a
behavior of reinforcement. Therefore, as mentioned in good choice to do material properties tests early, before
the paper, according to Saatcioglu and Grira1, to secure the test specimens are constructed.
the ductile behavior of a column under lateral loading, • As seen in Fig. 2 of the paper, diagonal bars are used
the grid bar with a welded joint should have at least 4% along the entire length of the beam and in the transverse
elongation capacity in tension. However, as observed hoops in the beam, which is more than the conventional
in Table 2 of the paper, no such information is present. cast-in-place specimen. Undoubtedly, such an
Moreover, mechanical properties of coupler splices and arrangement of shear reinforcement would change the
headed bars should be also included in the paper. behavior of the column—that is, the diagonal bars along
• It is indicated in the paper that axial load was not applied the length of the beam strengthen the fixing of bottom
to columns and that the performance of such reinforcing reinforcing bars, resulting in minimizing the slip at the
bar details is not affected by the axial load of columns. joint, as described in Fig. 2. Maybe it would be better to
The discussers do not agree with the authors on this point. use a specimen with only diagonal bars in the beam to
As is well known, the behavior of the beam-column make comparisons between behaviors of test specimens.
connection is more complicated with the existence of • In the paper, techniques of reinforcing bar welding,
axial load than with no axial load, and the steel band coupler splicing, and headed-bar anchorage are used
plates would expand around the column, which may for the beam-column connection; however, not much
have an unexpectedly adverse effect on the reinforcing more information is presented on the headed-bar
bar couplers. In addition, with the slenderness ratio of anchorage and reinforcing bar welding. In addition,
the column not exceeding 3 (2100/700 equates to 3), it when constructing the specimens, how are the couplers
can be defined as a short column; the existence of axial constructed to connect the reinforcing bars from the
load can cause compression failure to the column and beam and from the connection? It can be seen in Fig. 8(b)
make the behavior of the connection more complex. that reinforcing-bar slip occurred due to the loosened
Moreover, as mentioned in the recently pubished thread of the couplers, which indicated that there is a

ACI Structural Journal/March-April 2014 453


possibility of the couplers not working in practice.
Perhaps an alternative is to place the couplers outside
the beam-column connection to connect the reinforcing
bars of the beam, not in the connection region.
To sum up, the construction method of a prefabricated
reinforcing bar for beam-column connection by
reinforcing-bar welding, coupler splicing, and headed-bar
anchorage helps to accelerate the construction speed and is
a good choice for a precast concrete frame system. Because
the behavior of the beam-column connection is vital to the
entire structural system, much more experimental research
on such connections and structural frame systems with such
connections should be conducted to obtain enough reliable
proof for the proposed construction method, especially Fig. 15—Direct tension tests on three D25 bars with tag
focusing on welding reinforcement, which has been pointed welding joint.
out by the authors. The discussers are waiting with great
interest for the discussions in the announced follow-up paper.

REFERENCES
16. Priestley, M. J. N., “Overview of PRESSS Research Program,” PCI
Journal, July-Aug. 1991, pp. 50-57.
17. fib Bulletin 27, “Seismic Design of Precast Concrete Building
Structures, ” fib, Lausanne, Switzerland, 2003, pp. 1-2.
18. ACI Committee 374, “Guide for Testing Reinforced Concrete
Structural Elements under Slowly Applied Simulated Seismic Loads (ACI
374.2R-13),” American Concrete Institute, Farmington Hills, MI, 2013,
18 pp.

AUTHORS’ CLOSURE
The authors would like to thank the discussers for their Fig. 16—Simplified models for beam-column connections
interest in the paper and the informative discussion. The subjected to seismic loading.
authors’ response to the five comments is presented as
follows.
1. In the proposed prefabricated reinforcing bar construc- 2. As the discusser commented, when columns are
tion method (PRC method), tag welding is used for the subjected to high axial load, the axial load effect on the
connection between the transverse D13 bars and longitu- connection behavior may be undesirable and even critical.
dinal D25 and D22 bars. However, as shown in Fig. 5, the Thus, the overall behaviors of the beam-column connec-
amount of tag welding is relatively small, when compared tion specimens would be more realistic if axial load was
to the area of the longitudinal bars. Thus, the adverse effect applied to the columns. However, this test was performed
of tag welding on the longitudinal bars was expected to be to investigate the effect of the reinforcing bar details (such
minimal. As presented in the conclusion No. 5 of the paper, as the couplers, headed bars, band plates, and bar welding)
in this test, the bar tag welding did not have detrimental rather than the behavior of beam-column connections itself.
effects on the structural performance of the specimens. In the test specimens, the reinforcing bar details, except
The authors performed direct tension tests on the D25 bars the proposed details, are the same as the conventional
that had a tag welding joint to a transverse D13 bar at the ones. Thus, if the proposed details do not affect the overall
center (refer to Fig. 15). The results showed that two behavior of the specimens, the axial load effect on the overall
D25 bars were fractured away from the weld joints, but the behavior of the specimens should be the same as the effect
third specimen failed near the weld joint. Note that the elon- on beam-column connections with conventional reinforcing
gation capacities at rupture were much greater than 4%. bar details.
A material test for the steel used in the bar couplers The steel band plates are used to connect the beam rein-
was not performed. The yield and tensile strengths provided forcing bars to the column reinforcing bars for reinforcing
by the manufacturer were fy = 751 MPa (109 ksi) and fu = bar fabrication and erection, and to provide additional
760 MPa (110 ksi), which were much greater than those of bearing resistance for bar anchorage. Furthermore, as shown
the D22 and D25 bars. The couplers used in this test are in Fig. 3, conventional ties or hoops specified in current
commercial products, and the mechanical properties were design codes are used in the beam-column joints. Thus,
proved elsewhere. However, in this test, reinforcing bar slip the use of the steel band plates in addition to the conven-
occurred due to the loosened threads of the coupler in Speci- tional ties is not likely to affect the overall behavior of the
mens PRC1. Thus, attention should be paid to the coupler beam-column connections. Thus, the axial load effect on the
splice. proposed connection method is expected to be the same as
The headed bars used in the test were different from the the effect on conventional beam-column connections.
conventional one specified in the design code. As shown 3. There is a misunderstanding. As shown in Table 1 and
in Fig. 4, the beam flexural bars in the exterior joint were Fig. 3, the number of the top flexural bars was greater than
anchored to the steel band plate by nuts and washers. Because the number of the bottom bars.
the steel band plate provides additional bearing capacity for The test setup was planned to simulate the lateral move-
bar anchorage, the headed bars were expected to be safe. In ments of the interior and exterior beam-column connections
this test, failure did not occur in the headed bars. under seismic loading, as presented in Fig. 16. The test

454 ACI Structural Journal/March-April 2014


setup has been used in many previous tests of beam-column cyclic behaviors of RC1 and PRC1 was observed (compare
connections. the pinching in the cyclic curves presented in Fig. 7).
The material tests on the reinforcing bars used in the test 5. Please refer to the authors’ reply No. 1 regarding the bar
program were conducted after the cyclic load tests. Because welding, coupler splice, and headed bar anchorage.
the D25 bars used in this test are commercial products In the proposed method, the reinforcing bar cages of
and the mechanical properties were originally guaranteed columns and beams are fabricated in a fabrication shop
by the manufacturer, material tests were not performed separately. In the construction field, the column bar cage
before testing. However, in the material tests, the elonga- is erected first. Then beam bar cage is connected to the
tion at failure was proven to be unqualified. As the discussers column bar cage using the bar couplers, which are located
commented, a reinforcing bar tension test before the connec- at the column face. The location of the couplers was deter-
tion test is desirable and appropriate. mined considering the efficiency and economy in fabrication
4. The diagonal D13 bars in the beam are nonstructural and shipping. In a mockup test, the reinforcing bar fabri-
elements that are used only for reinforcing bar fabrication. cation of the PRC connection specimens was conducted
Therefore, conventional vertical stirrups are required for the without difficulty. In real construction fields, however, the
shear capacity of the beam. Further, as shown in Fig. 3, the PRC method may be a challenging construction method.
number of the diagonal bars is significantly less than that of Although the present study focused on verifying the struc-
the vertical stirrups. Thus, the effect of the diagonal bars on tural performance of the proposed beam-column connection
the overall behavior of the specimens was expected to be method, further improvement in the construction technique
minimal. As shown in Fig. 7, no notable difference in the is required for the practical use of the proposed method.

Disc. 110-S32/From the May-June 2013 ACI Structural Journal, p. 415.

Shear Strength of Reinforced Concrete Walls for Seismic Design of Low-Rise Housing. Paper by Julian
Carrillo and Sergio M. Alcocer
Discussion by Dun Wang and Xilin Lu
ACI member, PhD candidate, Research Institute of Structural Engineering and Disaster Reduction, School of Civil Engineering, Tongji University, Shanghi China; Professor,
Research Institute of Structural Engineering and Disaster Reduction, School of Civil Engineering, Tongji University.

It is known that various models have been proposed to • As noted by the authors, several existing proposed
estimate the shear strength of shear walls, such as those based models for estimating the shear strength of concrete
on the softened truss model initially developed by Mau and walls have limitations, which are described in the
Hsu (1987) and later modified by Guta (1996), the softened paper. Special attention should be paid to the fact that
strut-and-tie model proposed by Hwang et al. (2001), and the most of these proposed shear models are based on
UCSD shear model by Kowasky and Priestley (2000), which their individual experimental research and have their
was later modified by Krolicki (2011). The discussed paper own limitations to particular scopes, with different
presents a set of semi-empirical design equations that are influencing factors, loading protocols, and data-
capable of estimating the shear strength of walls for low-rise processing methods. Therefore, there is a need for a
housing, based on an experimental program of 39 isolated unified experimental research activity on shear strength
walls in cantilever, and discussion of previous experimental of walls to be launched worldwide due to the complex
studies by other researchers. As seen in the conclusions of mechanism of shear phenomena. In addition, research
this paper, the proposed equations qualify the efficiency on the shear strength of walls should also focus on a
factor of horizontal reinforcements by their contributions model based on theoretic analysis and then be verified
to the shear strength of walls in one- to two-story concrete by experimental results, such as the softened truss model
buildings in Latin America. The discussers would like to theory proposed by Mau and Hsu (1987) and modified
raise the following significant issues and suggestions: by Guta (1996), or the softened strut-and-tie model by
• As seen in Eq. (2), the concrete contribution to the shear Hwang et al. (2001); it should not be based merely on
strength of walls is associated to a1 and fc′, and a1 is experimental results.
related to the shear span ratio from Eq. (7), all of which • The paper includes wall specimens with web
are constant because the wall specimen is cast. However, reinforcement made of welded-wire mesh. As
with the increase of displacement, the cracks appear and mentioned by the authors, displacement ductility of
widen in concrete, resulting in reducing the effectiveness such a wall type may be limited by the small elongation
of the aggregate interlock shear resistance along the capacity of cold-drawn wire reinforcement; it can be
crack surface. Moreover, vertical reinforcements in the seen from Table 2 that the elongation is only 1.4% to
web wall may also contribute to the shear resistance, 1.9%. The discussers wonder whether it is allowed by
but none of these are included in the proposed model design codes to use such cold-drawn wire reinforcement
by the authors. The UCSD shear model seems to give with low elongation for web reinforcement of walls,
a relatively reasonable mechanism for the concrete especially for seismic design. In addition, Saatcioglu
contribution, which takes into account the effective shear and Grira (1999) recommended that, to secure the
area, the degradation of the shear resistance of concrete, ductile behavior of a column under lateral loading, the
the volumetric ratio of longitudinal reinforcement, and grid bar with a welded joint should have at least a 4%
the effect of the shear-span ratio. elongation capacity in tension. Moreover, it can be seen

ACI Structural Journal/March-April 2014 455


from Table 2 that the ratio of ultimate strength to yield load. Therefore, it is better to make out the condition for
strength for welded-wire mesh used in the paper is, at not considering the axial loading effect by quantitative
most, 1.16, which does not seem to have enough margin values. Moreover, as mentioned in the newly published
of strength for seismic design. ACI 374.2R-13 (ACI Committee 374 2013), gravity
• It is mentioned in the paper that wall reinforcement loads should be simulated during testing, whether their
of specimens was placed in a single layer at wall mid- effects are deemed important or not, because it would
thickness due to the thickness of 100 mm (4 in.) for result in a fast rate of lateral force strength degradation
one- and two-story buildings. However, no information with the existence of vertical loadings. Therefore, the
on the diameter of welded-wire mesh is provided in the contribution of vertical axial loads to the shear strength
paper, and it is not clear whether the concrete cover of walls should be included in Eq. (2), just like those
is too large to have a durability problem in practical proposed by Krolicki (2011).
projects with only one layer of reinforcement. That is to • Lastly, a clear definition for the squat wall needs to be
say, it is prone to cracks or shrinkage with single-layer made, and consensus should be reached on issues such
welded-wire mesh. In addition, in practice with a thicker as the influencing factors, the expression equation, and
wall section, there would be more than one layer of the mechanisms of each contribution to shear strength.
reinforcement in the wall where tie bars for connecting After that, comparisons between the previous proposal
the layers of reinforcement would be essential. So the models can be made to get a conclusive judgment for
contribution of tie bars to the shear strength of the wall the shear strength of the shear wall. Maybe there is
should also be taken into account. still a long way to go to determine the mechanisms of
• In Eq. (1) and (2), Aw is the area of wall concrete section each contribution to the shear strength of walls based
used to calculate the shear strength. In Eq. (6) and (8), on theoretic analysis, not merely on the experimental
lwtw is used. It is not mentioned whether the two indexes, research, whether the shear wall is a squat wall or
Aw and lwtw, are equal to each other. What is the value for flexural wall.
Aw if they are not the same? Maybe the calculation of
Aw should be clarified because there are different values REFERENCES
used in different models, such as those by Kassem ACI Committee 374, 2013, “Guide for Testing Reinforced Concrete
Structural Elements under Slowly Applied Simulated Seismic Loads (ACI
(2010) and Krolicki (2011). 374.2R-13),” American Concrete Institute, Farmington Hills, MI, 18 pp.
• As seen in Eq. (2), the web steel contribution to Gupta, A., 1996, “Behavior of High Strength Concrete Structural Walls,”
the shear strength of walls is limited to horizontal PhD thesis, Curtin University of Technology, Perth, Australia.
reinforcement and is associated with the yield strength Hwang, S. J.; Fang, W. H., Lee, H. J.; and Yu, H. W., 2001, “Analytical
of reinforcement, fyh, the volumetric ratios rh and hh, as Model for Predicting Shear Strength of Squat Walls,” Journal of Structural
Engineering, ASCE, V. 127, No. 1, pp. 43-50.
well as Aw, which has nothing to do with the longitudinal Kassem, W., and Elsheikh, A., 2010, “Estimation of Shear Strength of
steel reinforcement. However, as referred to by the Structural Shear Walls,” Journal of Structural Engineering, ASCE, V. 136,
authors in the paper, according to Barda et al. (1977), No. 10, Oct., pp. 1215-1224.
for a wall with hw/tw < 1 and with boundary elements, Kowalsky, M. J., and Priestley, M. J. N., 2000, “Improved Analytical
Model for Shear Strength of Circular Reinforced Concrete Columns in
horizontal reinforcement becomes less effective as Seismic Regions,” ACI Structural Journal, V. 97, No. 3, May-June, pp. 388-
compared to vertical reinforcement, particularly for 396.
walls with M/Vlw < 0.5, which indicates that vertical Krolicki, J.; Maffei, J. G.; and Calvi, M., 2011, “Shear Strength of
reinforcement contributes to the shear strength of the Reinforced Concrete Walls Subjected to Cyclic Loading,” Journal of
wall. Both of the cases— hw/tw < 1 and M/Vlw < 0.5—are Earthquake Engineering, V. 15, No. S1, pp. 30-71.
Mau, S. T., and Hsu, T. C., 1987, “Shear Behavior of Reinforced Concrete
all included in the scope of the proposed shear-strength Framed Wall Panels with Vertical Loads,” ACI Structural Journal, V. 84,
equations by the authors; that is, the proposed equations No. 3, May-June, pp. 228-234.
best predict peak shear strength of walls with M/Vlw Saatcioglu, M., and Grira, M., 1999, “Confinement of Reinforced
ratios less than or equal to 2.0. In addition, what is the Concrete Columns with Welded Reinforcement Grids,” ACI Structural
Journal, V. 96, No. 1, Jan.-Feb., pp. 29-39.
mechanism of the web horizontal reinforcement for the
contribution to shear resistance? Because it is calculated
as the products of fyh, rh, hh, and Aw, which is not similar AUTHORS’ CLOSURE
to that mentioned in Krolicki (2011)—that is, VS = The authors acknowledge the interest of discussers in
rttwhcrfy (as seen in detail in the reference). Moreover, the paper. Indeed, shear strength of concrete members is
the factor hh used for considering the yielding extent still an issue where the development of a unified approach
of horizontal web reinforcement is constant (0.8 for is needed. Studies, both experimental and analytical,
deformed bars and 0.7 for welded-wire mesh), which contribute to providing data, evidence, and reflections on the
does not reflect the effects of other influences such as phenomenon. Our research was aimed at developing simple
the imposed lateral drift, volumetric ratio of horizontal design tools that could be used in practice. The following are
reinforcement, axial loading, and yielding strength. comments based on discussers’ items:
• An assumption that the contribution of vertical axial 1. The model developed in our research recognizes that
stress to the shear strength of squat walls is unimportant factors hh,v, a1, and a2 do depend on wall ductility or wall
is made in the paper by the authors. Therefore, there drift. The model discussed in the paper is applicable to drifts
is no contribution of vertical axial loading effect in associated with peak strength (pp. 418 and 422). The model
Eq. (2). However, as a matter of fact, axial stresses developed also recognizes the contribution of vertical web
always exist whether it is increased or decreased by reinforcement to wall shear strength (pp. 420-422). In fact,
vertical acceleration or coupling between walls. In strains measured in the vertical web bars or steel wires during
addition, whether it is conservative, and to what extent, tests were mainly associated to the uniform distribution of
is not clear without the contribution of vertical axial inclined cracks on the wall web. It was confirmed that, as
specified by ACI 318,1 a minimum amount of web vertical

456 ACI Structural Journal/March-April 2014


steel should be computed using Eq. (5). On the other hand, wall shear strength. An equation to compute the vertical web
the accuracy and variables included in the UCSD shear steel ratio is included. The equation to calculate the wall
model are not within the scope of the paper. shear strength explicitly includes the horizontal web steel
2. The authors of the paper agree with the discussers ratio only. However, for this equation to be applicable, the
that there is a need for a unified approach for experimental vertical web steel ratio should be placed in the wall and
research on shear strength of walls to be launched worldwide computed using the proposed Eq. (5). Differences between
due to the lack of a consistent trend of the contribution of the proposed model and Krolicki model are not within
horizontal and vertical web reinforcement to wall peak the scope of the paper and, therefore, should be studied
shear strength (p. 418). As the discussers indicate, other in future research. In the case of the so-called “efficiency
methodologies for assessing shear behavior have been factor of horizontal web steel ratio,” as argued in Note 1,
developed, such as truss models and softened strut-and-tie the model acknowledges that hh depends on wall drift. The
models. Such models were not discussed in the paper and, model proposed in the paper is applicable to calculation of
therefore, should be studied in future research. maximum load-carrying capacity and, thus, the proposed
3. The authors agree with the discussers that steel equation is applicable to drifts associated with peak strength.
reinforcement to be used in seismic design applications must In addition, as discussed on p. 421 of the paper, values of hh
exhibit minimum elongation capacity. The purpose of using should be used when rhfyh is lower than 1.25 MPa (181 psi).
the welded-wire mesh available was to assess its adequacy as 7. The authors agree with the discussers that axial stress
reinforcement for seismic design applications (p. 415). The is important to accurately determine wall shear strength
authors have recommended that welded-wire mesh with low and deformation capacities. As discussed on p. 423 of the
elongation capacity should not be used for shear resistance. paper, for these box-type low-rise structures, axial stresses
4. Space limitations in the paper hindered the possibility on walls typically have small values. Therefore, assuming
of including details of the experimental program such as sv ≈ 0 is conservative for low-to-medium seismic hazard
the diameter of the bars and wires. Discussers are invited areas. A detailed analysis is warranted for structures located
to revisit the recommended references (Carrillo and in areas of high seismic hazard.
Alcocer 2012, 2013). Regarding the use of one curtain of 8. The authors agree with the discussers that more work
reinforcement within the wall web, Section 21.9.2.2 of ACI needs to be done to better differentiate shear-dominated and
318-11 also allows its use when the factored shear force is flexural-governed walls, as well as on the development of
lower than a predefined value, as is the common case in low- a comprehensive and unified theory on shear strength and
rise concrete housing. deformation capacities. These issues and the comparisons
5. The calculation of Aw is clearly specified after Eq. (2) between the previous proposal models are not within the
as the gross area of wall concrete bounded by wall thickness scope of the paper and should be the topic of a new study.
and wall length (Aw = twlw).
6. As indicated above, the model reported in the paper
considers the contribution of vertical web reinforcement to

Disc. 110-S39/From the May-June 2013 ACI Structural Journal, p. 491.

Performance of AASHTO-Type Bridge Model Prestressed with Carbon Fiber-Reinforced Polymer


Reinforcement. Paper by Nabil Grace, Kenichi Ushijima, Vasant Matsagar, and Chenglin Wu
Discussion by José R. Martí-Vargas
Associate Professor, ICITECH, Institute of Concrete Science and Technology, Universitat Politècnica de València, Valencia, Spain.

The discussed paper presents an interesting experimental the 12,141 mm (39 ft 10 in.) span length are considered,
study on the design philosophy, construction techniques the beam is related to the AASHTO Type VI I-beam, which
employed, and flexural performance of both an AASHTO is 1830 mm (72 in.) depth and is often used for span lengths
I-beam and a bridge model reinforced and prestressed with ranging from 36 to 45 m (118 ft 1.32 in. to 147 ft 7.65 in.)
carbon fiber composite cable (CFCC) strands. from the corresponding width and span length obtained as
The authors should be complimented for producing a 3.6 × 502 mm = 1807 mm (71.14 in.) and 3.6 × 12.141 m
detailed paper with comprehensive information. This is = 43.7 m (143 ft 4.77 in.), respectively. On the other hand,
acknowledged by the discusser, who would like to offer the the 95 mm (3.75 in.) width of the beam web results in 3.6 ×
following comments and questions for their consideration 95 mm = 342 mm (13.5 in.), which seems to be a broad width
and response: to be related to the 205 mm (8.1 in.) width of the AASHTO
1. A scale factor of 1:3.6 was used for both the AASHTO- Type V and Type VI I-beams. What were the criteria applied
Type control I-beam with a span of 12,141 mm (39 ft 10 in.) to choose the scale factor of 1:3.6 and the width of the beam
and 502 mm (19.75 in.) deep and the 2500 mm (98.75 in.) web? It is worth noting that different effects on length, cross
width bridge model made up of five such beams. Regarding section, and second area moment can result in an identical
the cross section of the beam, it seems that the scale factor scale factor.
was taken in relation to the AASHTO Type V or Type VI 2. Beam spacing in a bridge depends on several factors such
I-beam when based on the 203 mm (8 in.) width of the bottom as span length, concrete strength, loads, and environment,
flange; –3.6 × 203 = 731 mm (28.77 in.) is next to 712 mm among others. What was the criterion used to establish a
(28 in.), which is the actual bottom flange of AASHTO beam spacing of 502 mm (19.75 in.) in the bridge model?
Type V and Type VI I-beam. In addition, and according to Regarding the average 28-day concrete compressive strength
Gerges and Gergess,14 if the 502 mm (19.75 in.) depth or in the six AASHTO I-beams of 44.82 MPa (6500 psi),

ACI Structural Journal/March-April 2014 457


502 mm beam spacing is approximately 1.13 to 1.45— of 13 mm Prestressing Strand,” Structural Engineering and Mechanics,
obtained as (3.6 × 502)/1600 and (3.6 × 502)/1250—times V. 26, No. 2, 2007, pp. 211-229.
20. Martí-Vargas, J. R.; Serna, P.; Navarro-Gregori, J.; and Pallarés,
the maximum beam spacing for 43.7 m (143 ft 4.5 in.) span L., “Bond of 13 mm Prestressing Steel Strands in Pretensioned Concrete
length when taking as a reference the charts provided by Members,” Engineering Structures, V. 41, 2012, pp. 403-412.
Gerges and Gergess14 for optimizing girder size and spacing. 21. Martí-Vargas, J. R.; Arbeláez, C. A.; Serna-Ros, P; and Castro-
3. Based on the allowable concrete stresses at prestress Bugallo, C., “Reliability of Transfer Length Estimation from Strand End
Slip,” ACI Structural Journal, V. 104, No. 4, July-Aug. 2007, pp. 487-494.
transfer, the maximum number of 15.2 mm (0.6 in.) 22. Martí-Vargas, J. R.; Serna, P.; and Hale W. M., “Strand Bond
diameter 7-wire steel strands that can be accommodated in Performance in Prestressed Concrete Accounting for Bond Slip,”
an AASHTO Type VI I-beam with a 43.7 m (143 ft 4.5 in.) Engineering Structures, V. 51, 2013, pp. 236-244.
span length is 45 to 55.14 The authors used 10 longitudinal 23. Martí-Vargas, J. R., and Hale, W. M., “Predicting Strand Transfer
CFCC strands per beam: seven non-prestressed and three Length in Pretensioned Concrete: Eurocode versus North American
Practice,” Journal of Bridge Engineering, ASCE, 2012, DOI:10.1061/
prestressed at 30% of their average breaking load. It seems (ASCE)BE.1943-5592.0000456.
the concrete stresses after prestress transfer were too small 24. Martí-Vargas, J. R.; Serna, P.; Navarro-Gregori, J.; and Bonet, J. L.,
to be representative of pretensioned concrete structures “Effects of Concrete Composition on Transmission Length of Prestressing
when compared with current practice (prestressing steel Strands,” Construction and Buiding Materials, V. 27, 2012, pp. 350-356.
25. Caro, L. A.; Martí-Vargas, J. R.; and Serna, P., “Time-Dependent
strands prestressed at 75%). This can explain the fewer Evolution of Strand Transfer Length in Pretensioned Prestressed Concrete
prestress losses obtained as compared with conventional Members,” Mechanics of Time-Dependent Materials, V. 17, No. 4, Nov.
steel prestressed members.15,16 2013, pp. 501-527, DOI: 10.1007/s11043-012-9200-2.
4. Apart from the prestressing force being small according 26. Martí-Vargas, J. R.; Ferri, F. J.; and Yepes, V., “Prediction of the
to the number of CFCC prestressed strands and the used Transfer Length of Prestressing Strands with Neural Networks,” Computers
and Concrete, V. 2, No. 2, 2013, pp. 187-209.
prestress level, a shorter transfer length is obtained. Transfer 27. CEN, “Eurocode 2: Design of Concrete Structures — Part 1-1:
length can be measured17,18 and/or predicted from different General Rules and Rules for Buildings,” European standard EN 1992-1-
equations based on equilibrium of forces19,20 or strand 1:2004:E, Comité Européen de Normalisation, Brussels, Belgium, 2004.
slips.21,22 As transfer length provisions differ for distinct 28. Martí-Vargas, J. R.; Arbeláez, C. A.; Serna-Ros, P.; Fernández-Prada,
M. A.; and Miguel-Sosa, P. F., “Transfer and Development Lengths of
codes and researchers,19,23 and no consensus has been Concentrically Prestressed Concrete,” PCI Journal, V. 51, No. 5, Sept.-Oct.
reached as to the main parameters to be considered,24-26 it 2006, pp. 74-85.
would be interesting to detail the transfer length provisions 29. Martí-Vargas, J. R.; García-Taengua, E.; and Serna, P., “Influence
used by the authors to design the prestressed beams. Do the of Concrete Composition on Anchorage Bond Behavior of Prestressing
authors have any information on measured transfer length? Reinforcement,” Construction and Buiding Materials, V. 48, 2013,
pp. 1156-1164.
5. The concrete stresses at prestress transfer can be
assumed with a linear distribution beyond the dispersion
AUTHORS’ CLOSURE
length,27 which is longer than transfer length and depends
The authors would like to thank the discusser for his
on beam depth and tendon position, among other factors.
interest, insightful discussion, and thoughtful observations
According to the scale factor and the beam depth, what
on the presented research paper. The beam cross section of
were both the dispersion and transfer lengths in relation to
the bridge model was inspired by the Taylor Bridge built
the beam length? In addition, transfer length remains within
over Assiniboine River in the Parish of Headingley, Mani-
the development length of prestressing strands.12,28,29 It
toba, Canada. The five-span AASHTO-type bridge with
is worth remarking that the development length equation
fiber-reinforced polymer (FRP) was built in 1997 and has
in AASHTO1 includes a 1.6 multiplier factor that is used
a span length of 31.25 m (102.53 ft) with a depth of 1.8 m
to avoid bond failure caused by inadequately developed
(5.9 ft) and a 200 mm (7.87 in.) thick composite deck.3 The
lengths in structural members whose depth is greater
prestressed beams for the Taylor Bridge are spaced at 1.8 m
than 610 mm (24 in.) because of high shear effects. The
(5.9 ft) with a bottom flange width of 660 mm (25.98 in.). The
tested beams were 502 mm in depth (19.75 in.)—lesser
span-depth ratio of the tested control beam and bridge model
than 610 mm (24 in.)—but the loading conditions regarding
beams were based on the AASHTO1 Table 2.5.2.6.3.1 tradi-
strand development length would differ in the full-scale
tional minimum depths. For a simple span length of 12.141 m
member, as the resulting depth of 3.6 × 502 mm = 1807 mm
(39.833 ft), the composite depth required for a prestressed
(71.14 in.) is greater than 610 mm (24 in.).
concrete precast I-beam is 0.045 times the span’s length—
that is, 0.045 × 12.5 m = 0.563 m (1.847 ft). The modeled
REFERENCES
14. Gerges, N., and Gergess, A. N., “Implication of Increased Live Loads beam cross section had an approximate scale factor of
on the Design of Precast Concrete Bridge Girders,” PCI Journal, V. 57, 1/3.6 for an AASHTO Type IV beam. The true scale model
No. 4, Fall 2012, pp. 78-95. of the AASHTO Type IV beam has narrower web and depth
15. Caro, L. A.; Martí-Vargas, J. R.; and Serna, P., “Prestress Losses and it would have posed difficulties in placing concrete,
Evaluation in Prestressed Concrete Prismatic Specimens,” Engineering
Structures, V. 48, 2013, pp. 704-715.
vibrating, and compacting; therefore, the web width and
16. Martí-Vargas, J. R.; Serna-Ros, P.; Arbeláez, C. A.; and Rigueira- depth were adjusted for the beam models based on the
Victor, J. W., “Bond Behaviour of Self-Compacting Concrete in constructibility in the laboratory setting.30 A similar adjust-
Transmission and Anchorage,” Materiales de Construcción, V. 56, No. 284, ment was made in bottom flange width to accommodate
2006, pp. 27-42. minimum spacing of the carbon fiber-reinforced polymer
17. Martí-Vargas, J. R.; Serna-Ros, P.; Fernández-Prada, M. A.; Miguel-
Sosa, P. F.; and Arbeláez, C. A., “Test Method for Determination of the (CFRP) reinforcement and clear cover requirements. All
Transmission and Anchorage Lengths in Prestressed Reinforcement,” such modifications done in the AASHTO Type IV beam for
Magazine of Concrete Research, V. 58, No. 1, 2006, pp. 21-29. suiting laboratory setup are clear from the dimensions given
18. Martí-Vargas, J. R.; Caro, L.; and Serna, P., “Experimental Technique in Fig. 1 to 3. Because of these modifications, the presented
for Measuring the Long-Term Transfer Length in Prestressed Concrete,”
Strain, V. 49, 2013, pp. 125-134.
beam cross section is not exactly a true scaled version of
19. Martí-Vargas, J. R.; Arbeláez, C. A.; Serna-Ros, P.; Navarro-Gregori, the AASHTO Type IV beam in terms of geometric proper-
J.; and Pallarés-Rubio, L., “Analytical Model for Transfer Length Prediction ties; nevertheless, the cross section does closely represent it.

458 ACI Structural Journal/March-April 2014


Also, it is worth mentioning that the State Departments of and developed new equations to predict transfer length.
Transportation (DOTs) do modify, to certain extent, the stan- However, for the investigation reported in the manuscript, the
dard AASHTO type prestressing beam shapes depending authors emphasized the flexural behavior of the AASHTO
on the local needs (Michigan 1800 girder, Wisconsin Type IV beams prestressed with the CFRP tendons.
type 70”, PennDOT 28” beams, and so on).31 Similarly, The significance of the presented experimental study
the scaled spacing of the bridge model was based on the has been to explain the design philosophy as per the ACI
AASHTO1 Table 4.6.2.2.2.b.1 limitations to validate load 440.4R-047 design guidelines and the Unified Design
distribution factors. Approach,6 the construction techniques employed, and
In the traditional design approach of the prestressed the flexural performance of an AASHTO I-beam and
concrete bridges, the emphasis is given more on cost opti- bridge model reinforced and prestressed with the CFRP
mization, and hence the designers try to use the full capacity strands. Moreover, the test results reported herein also
of the prestressed concrete beams. The mentioned reinforce- aid in validation of the results of analytical and numerical
ment ratio of 0.46% is based on the strain compatibility studies subsequently undertaken in the future, with lucidly
design approach for bonded prestressed and nonprestressed provided geometric and material details along with various
CFRP tendons arranged vertically.6 It may be noted that, response quantities.
unlike the balanced ratio for steel, the balanced ratio for
FRP bars/tendons does not signify yielding; rather, it signi- REFERENCES
fies failure/rupture of the bars. Also, substantial research 30. Wu, C., “Performance of Concrete I-Beam Bridge Prestressed with
CFCC Reinforcement,” master’s thesis, Lawrence Technological University
has been accomplished on the use of steel wire strands for (LTU), Southfield, MI, 2009.
the prestressing operations; therefore, the process has been 31. Pennsylvania Department of Transportation (PennDOT), Bridge
industrialized with a higher level of confidence and duly Design Drawings, http://www.dot.state.pa.us/Internet/BQADStandards.
specified guidelines. However, no such guidelines have nsf/home.
yet been evolved on the use of CFRP reinforcements as 32. Grace, N. F., “Transfer Length of CFRP/CFCC Strands for Double-T
Girders.” PCI Journal, V. 45, No. 5, 2000, pp. 110-126.
prestressing strands while conducting this research. There- 33. Grace, N. F.; Enomoto, T.; Abdel-Mohti, A.; Tokal, Y.; and
fore, suitably, the prestressing force level was chosen as 30% Puravankara, S., “Flexural Behavior of Precast Concrete Box Beams Post-
of the average measured breaking load of strands. Tensioned with Unbonded, Carbon-Fiber-Composite Cables,” PCI Journal,
Grace32 and Grace et al.33 investigated transfer length of the V. 53, No. 4, 2008, pp. 62-82.
CFRP reinforcement used in prestressed concrete box beams

ACI Structural Journal/March-April 2014 459


IN ACI MATERIALS JOURNAL
The American Concrete Institute also publishes the ACI Materials PDF versions of these papers are available for download at
Journal. This section presents brief synopses of papers appearing the ACI website, www.concrete.org, for a nominal fee.
in the current issue.

From the March-April 2014 issue

111-M11—Practical Approach for Assessing Lightweight factor. In general, a Vf of 0.5% is found to be an upper limit for the produc-
Aggregate Potential for Concrete Performance tion of SCC. A greater Vf can hinder the self-consolidating characteristics.
by Daniel Moreno, Patricia Martinez, and Mauricio Lopez For the assessment of the passing ability of FR-SCC, a modified J-ring
The properties and amount of lightweight aggregates used in a concrete setup containing six or eight bars instead of 16 bars is proposed. The
mixture can significantly influence its mechanical properties and density. passing ability of FR-SCC can be expressed as the ratio of diameter:height
at the center of the J-ring test. The passing ability can also be evaluated
Nevertheless, such influence cannot be accurately described and used in
using the L-box with a single bar instead of three bars.
practical application without an extensive experimental work. A practical
A superworkable concrete (SWC) requiring low consolidation energy
evaluation method for assessing the influence of lightweight aggregate on
can still be produced with a Vf of 0.75% when a viscosity-modifying
concrete properties is required to anticipate the performance of concrete in
admixture is incorporated to prevent segregation and blockage. For the
advance and choose the most suitable lightweight aggregate for a certain
tested fiber types, the average residual strength (ARS) in flexure is shown
structural application.
to increase with Vf. Steel fibers exhibited the highest ARS value.
In this paper, existing models are reviewed, generalized, and validated
to obtain a methodology for assessing the potential of the lightweight 111-M14—Effect of Misalignment on Pulloff Test Results:
aggregates to provide a specified concrete density, modulus of elasticity, Numerical and Experimental Assessments
and compressive strength. A practical evaluation methodology is proposed by Luc Courard, Benoît Bissonnette, Andrzej Garbacz, Alexander
and validated with four different lightweight aggregates, obtaining correla- Vaysburd, Kurt von Fay, Grzegorj Moczulski, and Maxim Morency
tions between measured and estimated density; modulus of elasticity; and
The successful application of a concrete repair system is often evalu-
compressive strengths of 93.4, 84.8, and 91.7%, respectively. Therefore,
ated through pulloff testing. For such in-place quality control (QC) testing,
this methodology allows practical and reliable comparison and selection of
the inherent risk of misalignment might affect the recorded value and
lightweight aggregates based on only one trial mixture.
eventually make a difference in the acceptance of the work. The issue of
eccentricity in pulloff testing has been ignored in field practice because it
111-M12—Axisymmetric Fiber Orientation Distribution of Short
is seen as an academic issue. This paper presents the results of a project
Straight Fiber in Fiber-Reinforced Concrete
intended to quantify the effect of misalignment on pulloff tensile strength
by Jun Xia and Kevin Mackie
evaluation and provide a basis for improving QC specifications if neces-
The anisotropic orientation distribution of short fibers in fiber-reinforced sary. The test program consisted first of an analytical evaluation of the
concrete and the impact on mechanical properties have been established in problem through two-dimensional finite element modeling simulations and,
past research. In this paper, the cast-flow induced anisotropic fiber distri- in a second phase, in laboratory experiments in which the test variables
bution was categorized as axisymmetric with respect to the cement paste were the misalignment angle (0, 2, and 4 degrees) and the coring depth
flow direction. The probabilistic spatial orientation for fibers is introduced (15 and 30 mm [0.6 and 1.2 in.]). It was found that calculations provide
using the beta distribution and uniformity parameters. Either theoretical or a conservative but realistic lower bound limit for evaluating the influence
approximate equations for the orientation factor and the probability density of misalignment upon pulloff test results: a 2-degree misalignment can
function of the crossing angle were derived for any arbitrary cut plane. The be expected to yield a pulloff strength reduction of 7 to 9%, respectively,
derived orientation factor equation can be used to quantify the degree of for 15 and 30 mm (0.6 and 1.2 in.) coring depths, and the corresponding
anisotropy via image analysis. This process is easier and more accurate decrease resulting from a 4-degree misalignment reaches between 13 and
because instead of detailed orientation information for every single fiber, 16%. From a practical standpoint, the results generated in this study indicate
only total fiber counts on the cut sections are needed. The probabilistic that when specifying a pulloff strength limit in the field, the value should
macromechanical properties, such as ultimate tensile strength, are estimated be increased (probable order of magnitude: 15%) to take into account the
based on the selected micromechanical model that defines the single fiber- potential reduction due to testing misalignment.
matrix interactions.
111-M15—Strength and Microstructure of Mortar Containing
111-M13—Mixture Design and Testing of Fiber-Reinforced Nanosilica at High Temperature
Self‑Consolidating Concrete by Rahel Kh. Ibrahim, R. Hamid, and M. R. Taha
by Kamal H. Khayat, Fodhil Kassimi, and Parviz Ghoddousi The effect of high temperature on the mechanical properties and micro-
An extensive testing program was undertaken to evaluate the applica- structure of nanosilica-incorporated mortars has been studied. Results show
that the incorporation of nanosilica increases both compressive and flex-
bility of a mixture-proportioning method proposed for shrinkage control
ural strengths significantly at both ambient and after a 2-hour exposure to
in fiber-reinforced concrete (FRC) in proportioning fiber-reinforced self-
752°F (400°C) temperatures; the strengths increase with the increase of
consolidating concrete (FR-SCC). The study also proposed test methods
nanosilica content. A significant decrease in strength was recorded for all
to evaluate workability of FR-SCC. The investigated fibers included
control and nanosilica-incorporated mortar specimens after a 2-hour expo-
polypropylene, steel, and hybrid fibers of different properties with fiber
sure to 1292°F (700°C) heat; however, nanosilica-incorporated specimens
lengths of 5 to 50 mm (0.20 to 1.97 in.). Fiber volume Vf ranged between
show higher residual strength than those without nanosilica. Microstruc-
0.25 and 0.75%. The study also aimed to determine the impact of fiber
tural analysis shows that nanosilica reduces the calcium hydroxide crystals
type and addition on key properties of the fresh and hardened concrete.
to produce more calcium silicate hydrate, the process that contributes to
Hardened properties included compressive, splitting tensile, and average
the strength and the residual strength of the material. In addition, the mate-
residual strengths. rial exhibits a stable structure state up to 842°F (450°C), while exposure to
Test results indicate that the proposed methodology to maintain constant higher temperatures results in a decomposition of hydration products.
mortar thickness over coarse aggregate and fibers can reduce any signifi-
cant drop in workability of FR-SCC, resulting from an increase in fiber

460 ACI Structural Journal/March-April 2014


111-M16—Prediction of Strength, Permeability, and Hydraulic simulating cement hydration, the evolution of ITZ is also simulated in this
Diffusivity of Ordinary Portland Cement Paste approach. Through simulations, the influences of several factors related
by B. Kondraivendhan and B. Bhattacharjee to concrete mixture proportion on ITZ are investigated. It is found that
In this investigation, the compressive strength, permeability, and ITZ thickness, as defined by the overall average porosity, can be reduced
hydraulic diffusivity of ordinary portland cement (OPC) paste have been by using finer aggregate, increasing aggregate volume fraction, reducing
predicted from the knowledge of mixture factors, such as water-cement water-cement ratio (w/c), or making the binder system finer. Following
ratio (w/c) and curing ages. The relationships for pore size distribution hydration, the ITZ thickness decreases continuously, but the difference of
porosity between ITZ and bulk paste keeps almost constant at mature ages.
(PSD) parameters, such as mean distribution radius r0.5, dispersion coef-
ficient d, and permeable porosity P, are functions of the w/c and curing
111-M19—C4AF Reactivity—Chemistry and Hydration of
ages, and are readily available in the literature. By using these relation-
Industrial Cement
ships, the compressive strength, permeability, and hydraulic diffusivity can
by Hugh Wang, Delia De Leon, and Hamid Farzam
be estimated from the w/c and curing age information. It is observed that the
predicted compressive strength closely matches the reported experimental The study described in this paper involved a close examination of
compressive strength. It is also observed that the estimated permeability one of the least-researched cement phases, 4CaO·Al2O3·Fe2O3 (C4AF
data also closely matches the reported experimental permeability data. or ferrite). The two ferrite materials used in this program were: 1)
The predicted hydraulic diffusivity follows a similar trend, as reported in extracted from an industrial clinker free of 3CaO·Al2O3 (C3A or alumi-
the literature. nate); and 2) a laboratory-synthesized C4AF compound. Both ferrite
materials were initially characterized using optical microscopy, X-ray
111-M17—Cracking Tendency of Lightweight Aggregate Bridge diffraction (XRD), and Raman spectroscopy. The hydration kinetics and
Deck Concrete reactivity of both materials were studied with a calorimetric technique.
by Benjamin E. Byard, Anton K. Schindler, and Robert W. Barnes Similar to aluminate, ferrite also demonstrated a strong early hydra-
tion rate. An adequate amount of sulfate (SO3) was needed to regulate
Early-age cracking in bridge decks is a severe problem that may reduce the ferrite hydration rate. Based on this work, an equivalent aluminate
the functional life of the structure. In this project, the effect of using light- content, C3Aeq, is proposed. The hydration of C3Aeq needs to be properly
weight aggregate on the cracking tendency of bridge deck concrete was controlled with the appropriate amount of sulfate. The concept, based
evaluated using testing frames that restrain movement due to volume on equivalent aluminate content for cement sulfate optimization, ensures
change effects from placement to cracking. Expanded shale, clay, and slate proper early hydration behavior from both C3A and C4AF and avoids
lightweight coarse and fine aggregates were used to produce internal curing, potential cement-admixture incompatibility problems, especially for
sand-lightweight, and all-lightweight concretes to compare their behavior cements containing little or no C3A content.
relative to a normalweight concrete in bridge deck applications.
Specimens were tested under temperature conditions that simulate 111-M20—Tailoring Hybrid Strain-Hardening Cementitious
summer and fall placement scenarios. Increasing the amount of pre-wetted Composites
lightweight aggregate in the concrete systematically decreased the density, by Alessandro P. Fantilli, Hirozo Mihashi, and Tomoya Nishiwaki
modulus of elasticity, and coefficient of thermal expansion of the concrete.
A class of fiber-reinforced concrete, commonly called strain-hard-
When compared to a normalweight concrete, the use of lightweight aggre-
ening cementitious composite (SHCC), can show very ductile behavior
gates in concrete effectively delays the occurrence of early-age cracking in
under tension. In the post-cracking stage, several cracks develop before
bridge deck applications.
complete failure, which occurs when tensile strains finally localize in one
of the formed cracks. To predict the mechanical performances of monofiber
111-M18—Multi-Aggregate Approach for Modeling Interfacial
SHCC, a cohesive model has been proposed. Such a model is used herein
Transition Zone in Concrete
to tailor hybrid SHCC, made with long and short fibers. By combining
by Hongyan Ma and Zongjin Li uniaxial tensile tests and the theoretical results of the model, the critical
Interfacial transition zone (ITZ) has long been of particular interest in value of the fiber-volume fraction can be evaluated. It should be considered
concrete technology. The limited sensitivity of experimental techniques as the minimum amount of long fibers that can lead to the formation of
makes it attractive to study ITZ using computer simulations. In this paper, multiple cracking and strain hardening under tensile actions. The aim of the
a multi-aggregate approach is proposed to simulate the formation of ITZ present research is to reduce such volume as much as possible, to improve
in concrete. In light of a modified status-oriented computer model for the workability, and to reduce the final cost of SHCC.

ACI Structural Journal/March-April 2014 461


REVIEWERS IN 2013
In 2013, the individuals listed on these pages served as technical reviewers of papers offered for publication
in ACI periodicals. A special “thank you” to them for their voluntary assistance in helping ACI maintain
the high quality of its publication program.

£aźniewska-Piekarczyk, Beata Ahmed, Ehab


Silesian University of Technology University of Sherbrooke
Rybnik, Poland Sherbrooke, QC, Canada
Aamidala, Hari Shankar Aidoo, John
Parsons Brinckerhoff Rose-Hulman Institute of Technology
Herndon, VA Terre Haute, IN
Abdalla, Hany Aire, Carlos
College of Technological Studies National Autonomous University of Mexico
Shuwaikh, Kuwait Mexico City, DF, Mexico
Abdel-Fattah, Hisham Akakin, Tumer
University of Sharjah Turkish Ready Mixed Concrete Association
Sharjah, United Arab Emirates Istanbul, Turkey
Abdelgader, Hakim Akbari, Reza
Tripoli University University of Tehran
Tripoli, Libyan Arab Jamahiriya Tehran, Tehran, Islamic Republic of Iran
Abdelrahman, Amr Akcay, Burcu
Heliopolis, Egypt Kocaeli University
Abo-Shadi, Nagi Kocaeli, Turkey
Robert Englekirk, Inc. Akiyama, Mitsuyoshi
Santa Ana, CA Waseda University
Abou-Zeid, Mohamed Tokyo, Japan
American University in Cairo Akkaya, Yilmaz
Cairo, Egypt Istanbul Technical University
Abu Yosef, Ali Maslak, Istanbul, Turkey
University of Texas at Austin Alam, A.K.M. Jahangir
Austin, TX Dhaka, Bangladesh
Achillopoulou, Dimitra Alam, M. Shahria
Democritus University of Thrace The University of British Columbia
Xanthi, Greece Kelowna, BC, Canada
Acun, Bora Alam, Mahbub
University of Houston Stamford University Bangladesh
Houston, TX Dhaka, Bangladesh
Adebar, Perry Al-Attar, Tareq
University of British Columbia University of Technology
Vancouver, BC, Canada Baghdad, Iraq
Adhikary, Bimal Al-Azzawi, Adel
Austin, TX Nahrain University
Agarwal, Pankaj Baghdad, Iraq
Indian Institute of Technology Roorkee Albahttiti, Mohammed
Roorkee, Uttarakhand, India Kansas State University
Aggelis, Dimitrios Manhattan, KS
University of Ioannina Albuquerque, Albéria
Ioannina, Ioannina, Greece Federal Center of Technological Education of Mato Grosso
Agustiningtyas, Rudi Cuiabá, Mato Grosso, Brazil
Ministry of Public Works Al-Chaar, Ghassan
Bandung, Indonesia U.S. Army Engineer Research and Development Center
Ahmadi, Jamal Champaign, IL
University of Science and Technology Alcocer, Sergio
Tehran, Tehran, Islamic Republic of Iran Institute of Engineering, UNAM
Ahmadi Nedushan, Behrooz Mexico City, DF, Mexico
Yazd University Aldajah, Saud
Yazd, Yazd, Islamic Republic of Iran United Arab Emirates University
Ahmed, Ayub Al-Ain, United Arab Emirates
Birla Institute of Technology and Science, Pilani Aldea, Corina-Maria
Pilani, Rajasthan, India St. Catharines, ON, Canada

462 ACI Structural Journal/March-April 2014


 REVIEWERS IN 2013
Alexander, Scott Aravinthan, Thiru
UMA Engineering, Ltd. University of Southern Queensland
Edmonton, AB, Canada Toowoomba, Queensland, Australia
Ali, Nisreen Arisoy, Bengi
Universiti Putra Malaysia Ege University
Serdang, Selangor, Malaysia Izmir, Turkey
Ali, Samia Aristizabal-Ochoa, Jose
University of Engineering and Technology, Lahore National University
Lahore, Punjab, Pakistan Medellin, Antioquia, Colombia
Aljewifi, Hana Armwood, Catherine
The University of Cergy-Pontoise University of Nebraska
Cergy-Pontosie Cedex, France Omaha, NE
Alkhairi, Fadi Asaad, Diler
Arabtech Jardaneh Gaziantep University
Amman, Jordan Gaziantep, Turkey
Alkhrdaji, Tarek Asamoto, Shingo
Structural Group Saitama University
Hanover, MD Saitama, Saitama, Japan
Allahdadi, Hamidreza Aslani, Farhad
Bangalore, India University of Technology, Sydney
Al-Manaseer, Akthem Sydney, New South Wales, Australia
San Jose State University Assaad, Joseph
San Jose, CA Notre Dame University
Al-Martini, Samer Beirut, Lebanon
University of Western Ontario Atamturktur, Sez
London, ON, Canada Clemson University
Al-Qaisy, Wissam Clemson, SC
Safe Mix Ready Concrete Athanasopoulou, Adamantia
Sharjah, United Arab Emirates AKMI Metropolitan College
Alqam, Maha Xalandri, Greece
The University of Jordan Attaalla, Sayed
Amman, Jordan ADR Engineering, Inc.
Alsiwat, Jaber Mission Hills, CA
Saudi Consulting Services Aviram, Ady
Riyadh, Saudi Arabia Simpson Gumpertz & Heger, Inc.
Aly, Aly Mousaad San Francisco, CA
Louisiana State University Awati, Mahesh
Baton Rouge, LA B.L.D.E.A.’s College of Engineering & Technology, Bijapur
Amani Dashlejeh, Asghar Bijapur, Karnataka, India
Tehran, Tehran, Islamic Republic of Iran Awida, Tarek
Anania, Laura KEO International Consultants
University of Catania Kuwait
Catania, Italy Ayano, Toshiki
Andersson, Ronny Okayama University
Hollviken, Sweden Okayama, Japan
Andrade, Jairo Aydin, Abdulkadir Cuneyt
Pontifical Catholic University of Rio Grande do Sul Ataturk University
Porto Alegre, RS, Brazil Erzurum, Turkey
Andriolo, Francisco Aydin, Ertug
Andriolo Ito Engenharia SC Ltda European University of Lefke
São Carlos, São Paulo, Brazil Nicosia, Turkey
Ansari, Abdul Aziz Aydın, Serdar
Quaid-e-Awam Engineering University Dokuz Eylul University
Nawabshah, Sindh, Pakistan I˙zmir, Turkey
Aqel, Mohammad Aykac, Sabahattin
King Saud University Gazi University, Faculty of Engineering and Architecture
Riyadh, Saudi Arabia Ankara, Turkey
Aragón, Sergio Azad, Abul
Holcim (Costa Rica) King Fahd University of Petroleum & Minerals
San Rafael, Alajuela, Costa Rica Dhahran, Saudi Arabia

ACI Structural Journal/March-April 2014 463


REVIEWERS IN 2013
Aziz, Omar Bayuaji, Ridho
University of Salahaddin Universiti Teknologi PETRONAS
Erbil, Iraq Perak, Malaysia
Babafemi, Adewumi Beddar, Miloud
Obafemi Awolowo University M’sila University
Ile-ife, Nigeria M’sila, Algeria
Bacinskas, Darius Bediako, Mark
Vilnius Gediminas Technical University CSIR - Building and Road Research Institute
Vilnius, Lithuania Kumasi, Ghana
Badger, Christian Bedirhanoglu, Idris
Bates Engineering, Inc. Dicle University
Lakewood, CO Diyarbakir, Turkey
Bae, Sungjin Behnoud, Ali
Bechtel Corporation Iran University of Science and Technology
Frederick, MD Tehran, Islamic Republic of Iran
Bai, Shaoliang Belleri, Andrea
Chongqing University University of Bergamo
Chonqqing, China Dalmine, Italy
Bai, Yongtao Benboudjema, Farid
Kyoto University Ecole normale supérieure de Cachan
Kyoto, Japan Cachan, France
Bakhshi, Mehdi Benliang, Liang
Arizona State University Shanghai, China
Tempe, AZ Bennett, Richard
Balaguru, P. The University of Tennessee
Rutgers, the State University of New Jersey Knoxville, TN
Piscataway, NJ Bentz, Dale
Balouch, Sana National Institute of Standards and Technology
University of Dundee Gaithersburg, MD
Dundee, UK Berry, Michael
Banibayat, Pouya Montana State University
New York, NY Bozeman, MT
Banić, Davor Beygi, Morteza
Civil Engineering Institute of Croatia Mazandaran University
Zagreb, Croatia Babol, Mazandaran, Islamic Republic of Iran
Baran, Eray Bharati, Raj
Atilim University National Institute of Technology Calicut
Ankara, Turkey Calicut, Kerala, India
Barbosa, Maria Bhargava, Kapilesh
Federal University of Juiz de Fora Bhabha Atomic Research Centre
Juiz De Fora, Brazil Mumbai, Maharashtra, India
Barroso de Aguiar, Jose Bhattacharjee, Bishwajit
University of Minho Indian Institute Of Technology, Delhi
Guimaraes, Portugal New Delhi, India
Bartos, Peter Bhatti, Abdul
University of the West of Scotland - Paisley National University of Sciences and Technology
Paisley, Scotland, UK Islamabad, Pakistan
Bashandy, Alaa Bilcik, Juraj
Menofiya University Slovak University of Technology in Bratislava
Shibin El-Kom, Egypt Bratislava, Slovakia
Batson, Gordon Bilek, Vlastimil
Clarkson University ZPSV a.s.
Potsdam, NY Brno, Czech Republic
Bayrak, Oguzhan Bilir, Turhan
University of Texas at Austin Bülent Ecevit University
Austin, TX Zonguldak, Turkey
Bayraktar, Alemdar Billah, Abu Hena
Trabzon, Turkey The University of British Columbia
Kelowna, BC, Canada

464 ACI Structural Journal/March-April 2014


 REVIEWERS IN 2013
Bimschas, Martin Burak, Burcu
Regensdorf, Switzerland Orta Dogu Teknik Universitesi
Bindiganavile, Vivek Ankara, Turkey
University of Alberta Burdette, Edwin
Edmonton, AB, Canada University Of Tennessee
Biolzi, Luigi Knoxville, TN
Politecnico di Milano Byard, Benjamin
Milan, Italy University of Tennessee at Chattanooga
Birely, Anna Chattanooga, TN
Texas A&M University Calixto, José
College Station, TX Universidade Federal de Minas Gerais (UFMG)
Birkle, Gerd Belo Horizonte, Brazil
Stantec Consulting Ltd. Campione, Giuseppe
Calgary, AB, Canada Universita Palermo
Bisschop, Jan Palermo, Italy
University of Oslo Canbay, Erdem
Oslo, Norway Middle East Technical University
Blair, Bruce Ankara, Cankaya, Turkey
Lafarge North America Cano Barrita, Prisciliano
Herndon, VA Instituto Politécnico Nacional/CIIDIR Unidad Oaxaca
Bobko, Christopher Oaxaca, Oaxaca, Mexico
North Carolina State University Canpolat, Fethullah
Raleigh, NC Yildiz Technical University
Bochicchio, Victor Istanbul, Turkey
Hamon Custodis Capozucca, Roberto
Somerville, NJ Faculty of Engineering
Bondar, Dali Ancona, Italy
Tehran, Islamic Republic of Iran Carino, Nicholas
Bondy, Kenneth Chagrin Falls, OH
Consulting Structural Engineer Carreira, Domingo
West Hills, CA Chicago, IL
Boshoff, William Carrillo, Julian
Stellenbosch University Universidad Militar Nueva Granada (UMNG)
Stellenbosch, Western Cape, South Africa Bogotá, D.C., Colombia
Boulfiza, Moh Carvalho, Alessandra
University of Saskatchewan Pontifical Catholic University of Goiás (PUC-Goiás)
Saskatoon, SK, Canada Goiânia, Goiás, Brazil
Bousias, Stathis Castles, Bryan
University of Patras Western Technologies, Inc.
Patras, Greece Phoenix, AZ
Bradberry, Timothy Castro, Javier
TxDOT Bridge Division Pontificia Universidad Catolica de Chile
Austin, TX Santiago, Chile
Braestrup, Mikael Catoia, Bruna
Ramboll Hannemann and Hojlund A/S Federal University of São Carlos (UFSCar)
Virum, Denmark São Carlos, São Paulo, Brazil
Brand, Alexander Cattaneo, Sara
University of Illinois at Urbana-Champaign Politecnico di Milano
Urbana, IL Milan, Italy
Brena, Sergio Cedolin, Luigi
University of Massachusetts Politecnico di Milano
Amherst, MA Milano, Italy
Brewe, Jared Cetisli, Fatih
CTLGroup Firat University
Skokie, IL Elazig, Turkey
Bui, Van Chai, Hwa Kian
BASF Admixtures, Inc. Tobishima Corporation
Cleveland, OH Noda, Chiba, Japan

ACI Structural Journal/March-April 2014 465


REVIEWERS IN 2013
Chakraborty, Arun Choi, Bong-Seob
Bengal Engineering and Science University Chungwoon University
Howrah, West Bengal, India Hongseong-Gun, Republic of Korea
Chang, Jeremy Choi, Eunsoo
Holmes Fire & Safety Hongik University
Christchurch, New Zealand Seoul, Republic of Korea
Chastre, Carlos Choi, Kyoung-Kyu
FCT/UNL Soongsil University
Lisbon, Portugal Seoul, Republic of Korea
Chaudhry, Asif Choi, Sejin
Geoscience Advance Research Laboratories University of California, Berkeley
Islamabad, Pakistan Albany, CA
Chawla, Komal Chompreda, Praveen
Indian Institute of Technology Kanpur Mahidol University
Kota, India Nakornpathom, Thailand
Chen, Chun-Tao Choong, Kokkeong
National Taiwan University of Science and Technology Universiti Sains Malaysia
Taipei, Taiwan Pulau Pinang, Seberang Perai Selatan, Malaysia
Chen, Hua-Peng Chorzepa, Migeum
The University of Greenwich Park Ridge, IL
Chatham, Kent, UK Chowdhury, Subrato
Chen, Qi Ultra Tech Cement, Ltd.
Boral Materials Technology Mumbai, Maharashtra, India
San Antonio, TX Chun, Sung-Chul
Chen, Shiming Mokpo National University
Tongji University Mooan-gun, Republic of Korea
Shanghai, China Chung, Deborah
Chen, Wei State University of New York
Wuhan University of Technology Buffalo, NY
Wuhan, HuBei, China Chung, Jae
Chen, Xi University of Florida
Shanghai, China Gainesville, FL
Cheng, Min-Yuan Chung, Lan
National Taiwan University of Science and Technology Dankook University
Taipei, Taiwan Seoul, Republic of Korea
Chi, Maochieh Claisse, Peter
WuFeng University Coventry University
Chiayi County, Taiwan Coventry, UK
Chiaia, Bernardino Cleary, Douglas
Politecnico di Torino Rowan University
Torino, Piedmont, Italy Glassboro, NJ
Chiang, Chih-Hung Clendenen, Joseph
Chaoyang University of Technology Pleasant Hill, IL
Wufong, Taichung, Taiwan Cobo, Alfonso
Chindaprasirt, Prinya Polytechnic University of Madrid
Khon Kaen University Madrid, Spain
Khon Kaen, Thailand Coelho, Jano
Chiorino, Mario AltoQi Informatica
Politecnico di Torino Florianopolis, Santa Catarina, Brazil
Torino, Italy Colombo, Matteo
Cho, Chang-Geun Politecnico di Milano
Chosun University Lecco, Italy
Gwangju, Republic of Korea Conley, Christopher
Cho, Jae-Yeol United States Military Academy
Seoul National University West Point, NY
Seoul, Republic of Korea Cordova, Carlos
Cho, Soon-Ho La Paz, Bolivia
Gwangju University Coronelli, Dario
Gwangju, Republic of Korea Politecnico di Milano
Milano, Italy

466 ACI Structural Journal/March-April 2014


 REVIEWERS IN 2013
Corral, Ramón Deng, Yaohua
Universidad Autónoma de Sinaloa Iowa State University
Los Mochis, Sinaloa, Mexico Ames, IA
Correal, Juan Detwiler, Rachel
Arcon Structural Engineers, Inc. Braun Intertec Corporation
Rancho Santa Margarita, CA Minneapolis, MN
Cortes, Douglas Devries, Richard
New Mexico State University Milwaukee School of Engineering
Las Cruces, NM Milwaukee, WI
Criswell, Marvin Dhanasekar, Manicka
Colorado State University Queensland University of Technology
Fort Collins, CO Brisbane, Queensland, Australia
Cueto, Jorge Dhonde, Hemant
Universidad de La Salle University of Houston
Bogota, Colombia Houston, TX
D. L., Venkatesh Babu Di Ludovico, Marco
Kumaraguru College of Technology University of Naples Federico II
Coimbatore, Tamil Nadu, India Naples, Italy
Dang, Canh Dias, WPS
University of Arkansas University of Moratuwa
Fayetteville, AR Moratuwa, Sri Lanka
Das, Sreekanta Diaz Loya, Eleazar
University of Windsor Louisiana Tech University
Windsor, ON, Canada Ruston, LA
Daye, Marwan Dilger, Walter
Intima International Co., Ltd. University of Calgary
Al-Khobar, Saudi Arabia Calgary, AB, Canada
de Brito, Jorge Ding, Yining
IST/TUL Dalian, China
Lisbon, Portugal Diniz, Sofia Maria
de Frutos, Jose Universidade Federal de Minas Gerais (UFMG)
Universidad Politecnica de Madrid Be lo Horizonte, Brazil
Madrid, Spain Do, Jeongyun
De Rooij, Mario Kunsan National University
TNO Kunsan, Jeonbuk, Republic of Korea
Delft, the Netherlands Dodd, Larry
Deb, Arghya Parsons Brinckerhoff
Indian Institute of Technology, Kharagpur Orange, CA
Kharagpur, West Bengal, India Dogan, Unal
Degtyarev, Vitaliy Istanbul Technical University
Columbia, SC Istanbul, Turkey
Dehn, Frank Dolan, Charles
University of Leipzig University of Wyoming
Leipzig, Germany Laramie, WY
Delalibera, Rodrigo Dongell, Jonathan
University of São Paulo Pebble Technologies
São Carlos, São Paulo, Brazil Scottsdale, AZ
Delatte, Norbert Dongxu, Li
Cleveland State University Nanjing University of Technology
Broadview Heights, OH Nanjing, Jiangsu, China
Demir, Ali Dontchev, Dimitar
Celal Bayar University University of Chemical Technology and Metallurgy
Manisa, Turkey Sofia, Bulgaria
Demir, Serhat Dragunsky, Boris
Blacksea Technical University Universal Construction Testing, Ltd.
Trabzon, Turkey Highland Park, IL
Den Uijl, Joop Du, Jinsheng
Delft University of Technology Beijing Jiao Tong University
Delft, the Netherlands Beijing, China

ACI Structural Journal/March-April 2014 467


REVIEWERS IN 2013
Du, Lianxiang El-Salakawy, Ehab
The University of Alabama at Birmingham University of Manitoba
Birmingham, AL Winnipeg, MB, Canada
Du, Yingang Elsayed, Tarek
Anglia Ruskin University Cairo, Egypt
Chelmsford, UK Emamy Farvashany, Firooz
Dubey, Ashish Perthpolis Pty, Ltd.
United States Gypsum Corp Perth, Western Australia, Australia
Libertyville, IL Eom, Tae-Sung
Dutta, Anjan Catholic University of Daegu
Indian Institute of Technology Guwahati Kyeongsan-si, Republic of Korea
Guwahati, Assam, India Erdem, T.
Dutton, John Izmir Institute of Technology
Edmonton, AB, Canada Izmir, Turkey
Eid, Rami Ergün, Ali
University of Sherbrooke Afyon Kocatepe University
Sherbrooke, QC, Canada Afyonkarahısar, Turkey
El Ragaby, Amr Evangelista, Luís
University of Manitoba Instituto Superior de Engenharia de Lisboa
Winnipeg, MB, Canada Lisbon, Portugal
Elamin, Anwar Fafitis, Apostolos
University of Nyala Arizona State University
Khartoum, Sudan Tempe, AZ
El-Ariss, Bilal Fanella, David
United Arab Emirates University Klein and Hoffman
Al Ain, United Arab Emirates Chicago, IL
Elbahar, Mohamed Fantilli, Alessandro
Ken Okamoto & Associates: Structural Engineering Politecnico di Torino
Rancho Santa Margarita, CA Torino, Italy
ElBatanouny, Mohamed Faraji, Mahdi
University of South Carolina Tehran, Islamic Republic of Iran
West Columbia, SC Fardis, Michael
Eldarwish, Aly Patras, Greece
Alexandria, Egypt Farghaly, Ahmed
El-Dash, Karim University of Sherbrooke
College of Technological Studies Sherbrooke, QC, Canada
Kuwait, Kuwait Faria, Duarte
El-Dieb, Amr Faculdade de Ciências e Tecnologia
Ain Shams University Caparica-Lisbon, Portugal
Abbasia, Cairo, Egypt Farrow, William
El-Hawary, Moetaz Lebanon, NJ
Kuwait Institute for Scientific Research Farzam, Masood
Safat, Kuwait Structural Engineering
Elkady, Hala Tabriz, Islamic Republic of Iran
National Research Centre (NRC) Fenollera, Maria
Giza, Egypt Universidade de Vigo
El-Maaddawy, Tamer Vigo, Spain
United Arab Emirates University Ferguson, Bruce
Al-Ain, Abu Dhabi, United Arab Emirates University of Georgia
Elmenshawi, Abdelsamie Athens, GA
University of Calgary Fernández Montes, David
Calgary, AB, Canada Madrid, Spain
El-Metwally, Salah Fernández Ruiz, Miguel
University of Hawaii at Manoa Ecole Polytechnique Federale De Lausanne
Honolulu, HI Lausanne, Vaud, Switzerland
Elnady, Mohamed Ferrara, Liberato
Mississauga, ON, Canada Politecnico di Milano
El-Refaie, Sameh Milan, Italy
El-Gama City, Mataria, Cairo, Egypt

468 ACI Structural Journal/March-April 2014


 REVIEWERS IN 2013
Ferrier, E. Gesund, Hans
Université Claude Bernard Lyon 1 University of Kentucky
Villerubanne, France Lexington, KY
Folino, Paula Ghafari, Nima
University of Buenos Aires Laval University
Buenos Aires, Argentina Quebec, QC, Canada
Foster, Stephen Giaccio, Craig
University of New South Wales AECOM
Sydney, New South Wales, Australia Melbourne, Victoria, Australia
Fouad, Fouad Giancaspro, James
University of Alabama at Birmingham University of Miami
Birmingham, AL Coral Gables, FL
Fradua, Martin Gilbert, Raymond
Feld, Kaminetzky & Cohen, P.C. The University of New South Wales
Jericho, NY Sydney, New South Wales, Australia
Francüois, Buyle-Bodin Girgin, Canan
University of Lille Yildiz Technical University
Villeneuve d’Ascq, France Istanbul, Turkey
Fuchs, Werner Gisario, Annamaria
University of Stuttgart Sapienza – Università di Roma
Stuttgart, Germany Rome, Italy
Furlong, Richard Gnaedinger, John
Austin, TX Con-Cure Corporation
G., Dhinakaran Chesterfield, MO
Sastra University Goel, Rajeev
Thanjavur, India CSIR - Central Road Research Institute
Gabrijel, Ivan Delhi, India
University of Zagreb Goel, Savita
Zagreb, Croatia Whitlock Dalrymple Poston & Associates
Galati, Nestore New York, NY
Structural Group, Inc. Gökçe, H. Süleyman
Elkridge, MD Gazi University
Gallegos Mejia, Luis Ankara, Turkey
Fundacion Padre Arrupe de El Salvador Gongxun, Wang
Soyapango, San Salvador, El Salvador Hunan University of Science and Technology
Gamble, William Xiangtan, China
University of Illinois at Urbana-Champaign Gonzales Garcia, Luis Alberto
Urbana, IL Lagging SA
Ganesan, N. Lima, Peru
National Institute of Technology González-Valle, Enrique
Calicut, India Madrid, Spain
Garboczi, Edward Grandić, Davor
National Institute of Standards and Technology University of Rijeka
Gaithersburg, MD Rijeka, Croatia
Garcez, Estela Grattan-Bellew, P.
Universidade Federal de Pelotas (UFPel) Materials & Petrographic Research G-B Inc.
Pelotas, RS, Brazil Ottawa, ON, Canada
Garcia-Taengua, Emilio Greene, Thomas
Universidad Politecnica de Valencia W. R. Grace & Co.
Valencia, Spain Houston, TX
Gardoni, Paolo Gribniak, Viktor
University of Illinois at Urbana-Champaign Vilnius Gediminas Technical University
Urbana, IL Vilnius, Lithuania
Gayarre, Fernando Gu, Xiang-Lin
Gijon, Spain Tongji University
Gayed, Ramez Shanghai, China
University of Calgary Guadagnini, Maurizio
Calgary, AB, Canada The University of Shefifeld
Sheffield, UK

ACI Structural Journal/March-April 2014 469


REVIEWERS IN 2013
Guan, Garfield Harris, Devin
University of Cambridge University of Virginia
Cambridge, UK Charlottesville, VA
Guimaraes, Giuseppe Harris, G Terry
Pontificia Universidade Católica do Rio de Janeiro Green Cove Springs, FL
Rio de Janeiro, Brazil Harris, Nathan
Güneyisi, Erhan Menlo Park, CA
Gaziantep University Hashemi, Shervin
Gaziantep, Turkey Seoul National University
Guo, Guohui Seoul, Republic of Korea
Overland Park, KS Hassan, Assem
Guo, Liping Toronto, ON, Canada
Southeast University Hassan, Maan
Nanjing, Jiangsu Province, China University of Technology
Guo, Zixiong Baghdad, Iraq
Huaqiao University Hassan, Wael
Quanzhou, Fujian, China University of California, Berkeley
Gupta, Ajay Berkeley, CA
M.B.M. Engineering College He, Xiaobing
Jodhpur, Rajasthan, India Chongqing Jiaotong University
Gupta, Pawan Chongqing, China
Post-Tensioning Institute He, Zhiqi
Phoenix, AZ Southeast University
Haach, Vladimir Nanjing, Jiangsu, China
University of São Paulo Heinzmann, Daniel
São Carlos, São Paulo, Brazil Lucerne University of Applied Sciences and Arts
Habbaba, Ahmad Horw, Switzerland
Technische Universität München Henry, Richard
Garching, Germany University of Auckland
Haddad, Gilbert Auckland, New Zealand
SNC-Lavalin, Inc. Himawan, Aris
St. Laurent, QC, Canada Singapore, Singapore
Haddadin, Laith Hindi, Riyadh
United Nations Saint Louis University
New York, NY St. Louis, MO
Hadi, Muhammad Hoehler, Matthew
University of Wollongong Encinitas, CA
Wollongong, New South Wales, Australia Holschemacher, Klaus
Hadje-Ghaffari, Hossain HTWK Leipzig
John A. Martin & Associates Leipzig, Germany
Los Angeles, CA Hong, Sung-Gul
Hagenberger, Michael Seoul National University
Valparaiso University Seoul, Republic of Korea
Valparaiso, IN Hooton, R. Doug
Hamid, Roszilah University of Toronto
Universiti Kebangsaan Malaysia Toronto, ON, Canada
Bangi, Selangor, Malaysia Hossain, Khandaker
Hamilton, Trey Ryerson University
University of Florida Toronto, ON, Canada
Gainesville, FL Hossain, Tanvir
Hansen, Will Louisiana State University
Brighton, MI Baton Rouge, LA
Harajli, Mohamed Hosseini, Ardalan
American University of Beirut Isfahan University of Technology (IUT)
Beirut, Lebanon Isfahan, Islamic Republic of Iran
Harbec, David Hoult, Neil
University of Sherbrooke Toronto, ON, Canada
Sherbrooke, QC, Canada Hover, Kenneth
Cornell University
Ithaca, NY

470 ACI Structural Journal/March-April 2014


 REVIEWERS IN 2013
Hrynyk, Trevor Ibrahim, Hisham
University of Toronto Buckland and Taylor, Ltd.
Toronto, ON, Canada North Vancouver, BC, Canada
Hsu, Thomas Ichinose, Toshikatsu
University of Houston Nagoya Institute of Technology
Houston, TX Nagoya, Japan
Hu, Jiong Ideker, Jason
Texas State University-San Marcos Oregon State University
San Marcos, TX Corvallis, OR
Hu, Nan Ince, Ragip
Tsinghua University Firat University
Beijing, China Elazig, Turkey
Huang, Chang-Wei İpek, Süleyman
Chung Yuan Christian University Gaziantep University
Chung Li, Taiwan Gaziantep, Turkey
Huang, Jianwei Irassar, Edgardo
Southern Illinois University Edwardsville Universidad Nacional del Centro de la Provincia de Buenos
Edwardsville, IL Aires (UNCPBA)
Huang, Xiaobao Olavarria, Buenos Aires, Argentina
GM-WFG/GM-N American Project Center Islam, Md.
Warren, MI Chittagong University of Engineering & Technology (CUET)
Huang, Yishuo Chittagong, Bangladesh
Chaoyang University of Technology Issa, Mohsen
Wufeng, Taichung, Taiwan University of Illinois at Chicago
Huang, Zhaohui Chicago, IL
Brunel University Izquierdo-Encarnación, Jose
London, UK Porticus
Hubbell, David San Juan, PR
Toronto, ON, Canada Jaari, Asaad
Hueste, Mary Beth Dera, Dubai, United Arab Emirates
Texas A&M University Jaeger, Thomas
College Station, TX Baenziger Partner AG
Hulsey, J. Chur, Switzerland
University of Alaska Jain, Mohit
Fairbanks, AK Nirma University
Hung, Chung-Chan Ahmedabad, Gujarat, India
University of Michigan Jain, Shashank
Ann Arbor, MI Delhi Technological University (DTU)
Husain, Mohamed New Delhi, India
Zagazig University Jamshidi, Masoud
Zagazig, Egypt Building and Housing Research Center (BHRC)
Husem, Metin Tehran, Islamic Republic of Iran
Karadeniz Technical University Jang, Seung Yup
Trabzon, Turkey Korea Railroad Research Institute
Hussain, Raja Uiwang, Gyongggi-do, Republic of Korea
King Saud University Jansen, Daniel
Riyadh, Saudi Arabia California Polytechnic State University
Hussein, Amgad San Luis Obispo, CA
Memorial University of Newfoundland Jau, Wen-Chen
St. John’s, NL, Canada National Chiao Tung University (NCTU)
Hwang, Chao-Lung Hsinchu, Taiwan
National Taiwan University of Science and Technology Jawaheri Zadeh, Hany
Taipei, Taiwan Miami, FL
Ibell, Tim Jayapalan, Amal
University of Bath Georgia Institute of Technology
Bath, UK Atlanta, GA
Ibrahim, Amer Jeng, Chyuan-Hwan
Baquba, Iraq National Chi Nan University-Taiwan
Puli, Nantou, Taiwan

ACI Structural Journal/March-April 2014 471


REVIEWERS IN 2013
Jensen, Elin Kenai, Said
Lawrence Technological University Université de Blida
Southfield, MI Blida, Algeria
Jeon, Se-Jin Kevern, John
Ajou University University of Missouri-Kansas City
Suwon-si, Gyeonggi-do, Republic of Korea Kansas City, MO
Jiang, Jiabiao Khan, Mohammad
W. R. Grace (Singapore) Pte. Ltd. King Saud University
Singapore Riyadh, Saudi Arabia
Jirsa, James Kheder, Ghazi
University of Texas at Austin University of Al Mustansiriya
Austin, TX Baghdad, Iraq
Jozić, Dražan Khennane, Amar
Universty of Split Australian Defense Force Academy, University of New South
Split, Croatia Wales (AFDA, UNSW)
Kabele, Petr Canberra, Australian Capital Territory, Australia
Czech Technical University in Prague Khuntia, Madh
Praha 6, Czech Republic DuKane Precast Inc.
Kagaya, Makoto Naperville, IL
Akita, Japan Kianoush, M. Reza
Kalkan, Ilker Ryerson University
Kırıkkale University Toronto, ON, Canada
Kırıkkale, Turkey Kilic, Sami
Kan, Yu-Cheng Bogazici University
Chaoyang University of Technology Istanbul, Turkey
Taichung County, Taiwan Kim, Jae Hong
Kanakubo, Toshiyuki UNIST
University of Tsukuba Ulsan, Republic of Korea
Tsukuba, Japan Kim, Jang Hoon
Kandasami, Siva Ajou University
Bristol, UK Suwon, Republic of Korea
Kang, Thomas Kim, Yail
Seoul National University University of Colorado Denver
Seoul, Republic of Korea Denver, CO
Kankam, Charles Kirgiz, Mehmet
Kwame Nkrumah University of Science & Technology Hacettepe University
Kumasi, Ghana Ankara, Turkey
Kansara, Kunal Kishi, Norimitsu
University of Bath Muroran Institute of Technology
Bath, UK Muroran, Japan
Kantarao, Velidandi Kisicek, Tomislav
Central Road Research Institute University of Zagreb
New Delhi, Delhi, India Zagreb, Croatia
Karahan, Okan Klemencic, Ronald
Erciyes University Magnusson Klemencic Associates
Erciyes, Turkey Seattle, WA
Karayannis, Christos Ko, Lesley Suz-Chung
Democritus University of Thrace Holcim Group Support, Ltd.
Xanthi, Greece Holderbank, AG, Switzerland
Karbasi Arani, Kamyar Koehler, Eric
University of Naples Federico II University of Texas at Austin
Napoli, Italy Austin, TX
Kawamura, Mitsunori Koenders, Eddy A. B.
Kanazawa, Ishikawa, Japan Delft University of Technology
Kazemi, Mohammad Delft, the Netherlands
Sharif University of Technology Kotsovos, Gerasimos
Tehran, Islamic Republic of Iran National Technical University of Athens
Kazemi, Sadegh Athens, Greece
University of Alberta Kotsovos, Michael
Edmonton, AB, Canada Athens, Greece

472 ACI Structural Journal/March-April 2014


 REVIEWERS IN 2013
Krem, Slamah Lee, Chadon
University of Waterloo Chung-Ang University
Waterloo, ON, Canada Ansung, Kyungki-do, Republic of Korea
Krstulovic-Opara, Neven Lee, Chung-Sheng
ExxonMobil Development Company University of California, San Diego
Houston, TX La Jolla, CA
Kumar, Pardeep Lee, Deuck Hang
University of California, Berkeley University of Seoul
Berkeley, CA Seoul, Republic of Korea
Kumar, Rakesh Lee, Douglas
Central Road Research Institute Douglas D. Lee & Associates
Delhi, India Fort Worth, TX
Kumar, Vinod Lee, Heui Hwang
Steel Authority of India Limited Arup
Ranchi, Jharkhand, India San Francisco, CA
Kunieda, Minoru Lee, Hung-Jen
Nagoya University National Yunlin University of Science and Technology
Nagoya, Japan Douliu, Yunlin, Taiwan
Kunnath, Sashi Lee, Jaeman
University of California, Davis Kyoto University
Davis, CA Kyoto, Japan
Kupwade-Patil, Kunal Lee, Jung-Yoon
Massachusetts Institute of Technology Sung Kyun Kwan University
Cambridge, MA Suwon, Republic of Korea
Kurtis, Kimberly Lee, Nam Ho
Georgia Institute of Technology SNC-Lavalin Nuclear Inc.
Atlanta, GA Oakville, ON, Canada
Kusbiantoro, Andri Lee, Seong-Cheol
Universiti Malaysia Pahang KEPCO International Graduate School (KINGS)
Gambang, Pahang, Malaysia Ulsan, Republic of Korea
Kwan, Albert Lee, Seung-Chang
The University of Hong Kong Samsung C&T Corporation
Hong Kong, China Seoul, Republic of Korea
Kwan, Wai Hoe Lee, Yoon-Si
Universiti Sains Malaysia Bradley University
Gelugor, Penang, Malaysia Peoria, IL
La Tegola, Antonio Lee, Young Hak
University of Lecce Seoul, Republic of Korea
Lecce, Italy Leiva Fernández, Carlos
LaFave, James University of Seville
University of Illinois at Urbana-Champaign Seville, Andalucia, Spain
Champaign, IL Leo Braxtan, Nicole
Lai, Jianzhong New York, NY
Nanjing University of Science and Technology Lequesne, Remy
Nanjing, Jiangsu, China University of Kansas
Larbi, Kacimi Madison, KS
University of Sciences and Technology of Oran Leutbecher, Torsten
Oran, Oran, Algeria Universität Kassel
Laskar, Aminul Kassel, Germany
National Institute of Technology Li, Fumin
Silchar, Assam, India China University of Mining and Technology
Laterza, Michelangelo Xuzhou, Jiangsu, China
University of Basilicata Li, Wei
Potenza, Italy Wenzhou University
Law, David Wenzhou, Zhejiang, China
RMIT University Lignola, Gian Piero
Melbourne, Victoria, Australia University of Naples
Lawler, John Naples, Italy
Wiss, Janney, Elstner Associates, Inc.
Northbrook, IL

ACI Structural Journal/March-April 2014 473


REVIEWERS IN 2013
Lima, Maria Cristina Machida, Atsuhiko
Federal University of Uberlândia Saitama University
Uberlandia, Minas Gerais, Brazil Saitama, Japan
Lin, Wei-Ting Macht, Jürgen
Institute of Nuclear Energy Research Kirchdorf, Austria
Taoyuan, Taiwan Maekawa, Koichi
Lin, Yiching University of Tokyo
National Chung-Hsing Univ Tokyo, Japan
Taichung, Taiwan Magliulo, Gennaro
Liu, Shuhua University of Naples Federico II
Wuhan University Naples, Italy
Wuhan, HuBei, China Magureanu, Cornelia
Liu, Yanbo Technical University of Cluj Napoca
Florida Department of Transportation-State Materials Office Cluj Napoca, Cluj, Romania
Gainesville, FL Mahboub, Kamyar
Lizarazo Marriaga, Juan University of Kentucky
Coventry University Lexington, KY
Coventry, UK Mahfouz, Ibrahim
Londhe, Rajesh Cairo, Egypt
Government College of Engineering Aurangabad Malik, Adnan
Aurangabad, Maharashtra, India University of New South Wales
Long, Adrian Sydney, Australia
Queens University Mancio, Mauricio
Belfast, Ireland University of California, Berkeley
Loo, Yew-Chaye Berkeley, CA
Gold Coast, Australia Mander, John
Loper, James Texas A&M University
Jacobs Facilities, Inc. College Station, TX
Arlington, VA Manso, Juan
Lopes, Anne University of Burgos
Furnas Centrais Elétricas S/A Burgos, Castilla-León, Spain
Aparecida De Goiania, Goias, Brazil Marikunte, Shashi
Lopes, Sergio Southern Illinois University
University of Coimbra Carbondale, IL
Coimbra, Portugal Martinelli, Enzo
López-Almansa, Francisco University of Salerno
Technical University of Catalonia Fisciano, Italy
Barcelona, Spain Martí-Vargas, José
Lounis, Zoubir Universitat Politècnica de València
National Research Council Valencia, Valencia, Spain
Ottawa, ON, Canada Maruyama, Ippei
Lubell, Adam Nagoya University
Read Jones Christoffersen, Ltd. Nagoya, Aichi, Japan
Vancouver, BC, Canada Maslehuddin, Mohammed
Ludovit, Nad King Fahd University of Petroleum and Minerals
Alfa 04 Dhahran, Saudi Arabia
Kosice, Slovakia Matamoros, Adolfo
Luo, Baifu University of Kansas
Harbin, China Lawrence, KS
Lushnikova, Nataliya Mathew, George
National University of Water Management and Nature
 Cochin University of Science and Technology
Resources Use Cochin, Kerala, India
Rivne, Ukraine Matsagar, Vasant
Ma, Zhongguo Lawrence Technological University
University of Tennessee Southfield, MI
Knoxville, TN Matta, Fabio
MacDougall, Colin University of South Carolina
Kingston, ON, Canada Columbia, SC

474 ACI Structural Journal/March-April 2014


 REVIEWERS IN 2013
Maximos, Hany Mezhov, Alexander
Pharos University in Alexandria Moscow State University of Civil Engineering
Alexandria, Egypt Moscow, Russian Federation
Mbessa, Michel Milestone, Neil
University of Yaoundé I - ENSP Callaghan Innovation
Yaoundé, Center, Cameroon Lower Hutt, New Zealand
McCabe, Steven Minelli, Fausto
Lawrence, KS University of Brescia
McCall, W. Brescia, Brescia, Italy
Concrete Engineering Consultants Mishra, Laxmi
Charlotte, NC Motilal Nehru National Institute of Technology Allahabad
McCarter, John Allahabad, Uttar Pradesh, India
Heriot Watt University Mlynarczyk, Alexandar
Edinburgh, UK Wiss, Janney, Elstner Associates, Inc.
McDonald, David Princeton Junction, NJ
USG Corporation Mohamad, Gihad
Libertyville, IL University of Extremo Sul Catarinense - UNESC
McLeod, Heather Alegrete, Rio Grande do Sul, Brazil
Kansas Department of Transportation Mohamed, Ashraf
Topeka, KS Alexandria University
Meda, Alberto Alexandria, Egypt
University of Bergamo Mohamed, Nayera
Bergamo, Italy Assiut University
Medallah, Khaled Assiut, Egypt
Saudi Aramco IKPMS Mohammed, Tarek
Al Khobar, Saudi Arabia University of Asia Pacific
Meddah, Seddik Dhaka, Bangladesh
Kingston University London Montejo, Luis
Kingston, London, UK North Carolina State University
Mehanny, Sameh Raleigh, NC
Cairo University Moradian, Masoud
Cairo, Egypt Oklahoma State University
Meinheit, Donald Stillwater, OK
Wiss, Janney, Elstner Associates, Inc. Moreno Júnior, Armando
Chicago, IL Unicamp
Meininger, Richard Campinas, São Paulo, Brazil
Turner-Fairbank Highway Research Center-FHWA Moretti, Marina
Columbia, MD University of Thessaly
Mejia, Luis Gonzalo Athens, Greece
LGM & Cia Moriconi, Giacomo
Medellin, Colombia Technical University of Marche
Melchers, Robert Ancona, Italy
The University of Newcastle Morley, Christopher
Newcastle, New South Wales, Australia Cambridge University
Melo, José Cambridge, UK
University of Aveiro Moser, Robert
Aveiro, Portugal US Army Engineer Research and Development Center
Meng, Tao Vicksburg, MS
Institution of Building Materials Mostofinejad, Davood
Hangzhou, Zhejiang, China Isfahan University of Technology
Menon, Devdas Isfahan, Isfahan, Islamic Republic of Iran
Indian Institute of Technology Motaref, Sarira
Chennai, Tamilnadu, India University of Connecticut
Mermerdaş, Kasım Storrs, CT
Hasan Kalyoncu University Mubin, Sajjad
Gaziantep, Turkey University of Engineering and Technology
Meshgin, Pania Lahore, Punjab, Pakistan
University of Colorado Boulder Mulaveesala, Ravibabu
Boulder, CO Indian Institute of Information Technology
Jabalpur, India

ACI Structural Journal/March-April 2014 475


REVIEWERS IN 2013
Munoz, Jose Nowak, Andrzej
Federal Highway Administration University of Nebraska-Lincoln
McLean, VA Lincoln, NE
Muttoni, Aurelio Nwaubani, Sunny Onyebuchi
Swiss Federal Institute of Technology Anglia Ruskin University
Lausanne, Switzerland Chelmsford, UK
Nabavi, Esrafil O’Connor, Arthur
Rezvanshahr, Guilan, Islamic Republic of Iran O C Engineering
Nafie, Amr Reno, NV
Cairo, Egypt Offenberg, Matthew
Naish, David Rinker Materials
California State University, Fullerton Orlando, FL
Fullerton, CA Oh, Byung
Najimi, Meysam Seoul National University
University of Nevada, Las Vegas Seoul, Republic of Korea
Las Vegas, NV Okeil, Ayman
Nakamura, Hikaru Louisiana State University
Nagoya University Baton Rouge, LA
Nagoya, Aichi, Japan Olanitori, Lekan
Narayanan, Pannirselvam The Federal University of Technology Akure
VIT University Akure, Ondo State, Nigeria
Vellore, Tamilnadu, India Orakcal, Kutay
Narayanan, Subramanian Bogazici University
Gaithersburg, MD Istanbul, Turkey
Nassif, Hani Orr, John
Rutgers, The State University of New Jersey University of Bath
Piscataway, NJ Bath, UK
Negrutiu, Camelia Orta, Luis
Technical University of Cluj Napoca Instituto Tecnológico y de Estudios Superiores de Monterrey
Cluj Napoca, Cluj, Romania (ITESM)
Neves, Luís Zapopan, Jalisco, Mexico
University of Coimbra Ortega, J.
Coimbra, Portugal University of Alacant
Ng, Ernesto Alacant, Alicante, Spain
Maveang, S.A. Otieno, Mike
Panama, Panama University of Cape Town
Ng, Pui Lam Cape Town, Western Cape, South Africa
The University of Hong Kong Otsuki, Nobuaki
Hong Kong Tokyo Institute of Technology
Nichols, John Tokyo, Japan
Texas A&M University Ousalem, Hassane
College Station, TX Takenaka Corporation - Research and Development Institute
Nimityongskul, Pichai Inzai, Chiba, Japan
Asian Institute of Technology Ozbay, Erdogan
Pathumthani, Thailand Iskenderun, Hatay, Turkey
Nkinamubanzi, Pierre-Claver Ozden, Sevket
Institute for Research in Construction Kocaeli Universitesi
Ottawa, ON, Canada Kocaeli, Turkey
Nokken, Michelle Ozturan, Turan
Concordia University Bogazici University
Montreal, QC, Canada Istanbul, Turkey
Noor, Munaz Ozturk, Ali
Bangladesh University of Engineering and Technology Dokuz Eylul University
Dhaka, Bangladesh Izmir, Buca, Turkey
Noshiravani, Talayeh P S, Ambily
École polytechnique fédérale de Lausanne (EPFL) Council of Scientific & Industrial Research (CSIR)
Lausanne, Switzerland Chennai, Tamilnadu, India
Novak, Lawrence Pacheco, Alexandre
Portland Cement Association Universidade Federal do Rio Grande do Sul (UFRGS)
Skokie, IL Porto Alegre, Rio Grande do Sul, Brazil

476 ACI Structural Journal/March-April 2014


 REVIEWERS IN 2013
Page, Adrian Pirayeh Gar, Shobeir
University of Newcastle Houston, TX
Newcastle, New South Wales, Australia Polak, Marianna
Palmisano, Fabrizio University of Waterloo
Politecnico di Bari Waterloo, ON, Canada
Bari, Italy Potter, William
Pan, Wang Fook Florida Department of Transportation
SEGi University Tallahassee, FL
Petaling Jaya, Selangor, Malaysia Pourazin, Khashaiar
Pantazopoulou, Stavroula International Institute of Earthquake Engineering
 and
Demokritus University of Thrace Seismology (IIEES)
Xanthi, Greece Tehran, Islamic Republic of Iran
Pape, Torill Prakash, M. N.
University of Newcastle J.N.N. College of Engineering
Callaghan, New South Wales, Australia Shimoga, Karnataka, India
Parghi, Anant Prasittisopin, Lapyote
The University of British Columbia Oregon State University
Kelowna, BC, Canada Corvallis, OR
Park, Honggun Proske, Tilo
Seoul National University Technische Universität Darmstadt
Seoul, Republic of Korea Darmstadt, Germany
Parsekian, Guilherme Prota, Andrea
Federal University of São Carlos University of Naples
São Carlos, São Paulo, Brazil Naples, Italy
Patel, Raj Provis, John
CeraTech, Inc. University of Melbourne
Baltimore, MD Victoria, Australia
Pauletta, Margherita Prusinski, Jan
University of Udine Slag Cement Association
Tavagnacco, Udine, Italy Sugar Land, TX
Paulotto, Carlo Pujol, Santiago
Acciona S.A. Berkeley, CA
Alcobendas, Spain Puthenpurayil Thankappan, Santhosh
Pavlikova, Milena Granite Construction Company
CTU in Prague Abu Dhabi, United Arab Emirates
Prague, Czech Republic Putra Jaya, Ramadhansyah
Pellegrino, Carlo Universiti Teknologi Malaysia
University of Padova Skudai, Malaysia
Padova, Italy Qasrawi, Hisham
Peng, Cao The Hashemite University
Harbin Institute of Technology Zarqa, Jordan
Harbin, Heilongjiang, China Qiu, Bin
Peng, Jianxin Xi’an University of Architecture & Technology
Institute of Bridge Engineering Xi’an, China
Changsha, Hunan, China Rafi, Muhammad
Perez, Gustavo NED University of Engineering and Technology
Universidad Nacional de Tucumán Karachi, Sindh, Pakistan
Yerba Buena, Tucumán, Argentina Ragueneau, Frederic
Perez Caldentey, Alejandro Ecole normale supérieure de Cachan (ENS Cachan)
Universidad Politécnica de Madrid Cachan, France
Madrid, Madrid, Spain Rahal, Khaldoun
Persson, Bertil Kuwait University
Bara, Sweden Safat, Kuwait
Pessiki, Stephen Rajamane, N. P.
Lehigh University SRM University
Bethlehem, PA Kattankulathur, Tamil Nadu, India
Phillippi, Don Ramamurthy, K.
Diamond Pacific Indian Institute of Technology Madras
Rancho Cucamonga, CA Chennai, Tamilnadu, India

ACI Structural Journal/March-April 2014 477


REVIEWERS IN 2013
Ramaswamy, Ananth Rodrigues, Conrado
Indian Institute of Science Federal Centre for Technological Education in Minas Gerais
Bangalore, Karnataka, India (CEFET-MG)
Ramos, António Belo Horizonte, Minas Gerais, Brazil
Faculdade de Ciências e Tecnologia Rodrigues, Publio
Monte de Caparica, Portugal LPE Engenharia e Consultoria
Rangan, Vijaya São Paulo, Brazil
Curtin University of Technology Rodriguez, Mario
Perth, Western Australia, Australia National University of Mexico
Rao, Hanchate Mexico City, DF, Mexico
Jawaharlal Nehru Technological University
 College of Roh, Hwasung
Engineering Chonbuk National University
Anantapur, India Jeonju, Republic of Korea
Rao, Sarella Rosenboom, Owen
National Institute of Technology Hong Kong Polytechnic University
Warangal, Andhra Pradesh, India Hung Hom, Kowloon, Hong Kong
Raoof, Mohammed Rteil, Ahmad
Loughborough University University of British Columbia
Loughborough, UK Kelowna, BC, Canada
Rautenberg, Jeffrey Russell, Henry
Wiss, Janney, Elstner Associates, Inc. Henry G. Russell, Inc.
Emeryville, CA Glenview, IL
Ray, Indrajit S. A., Jaffer Sathik
Purdue University Calumet CSIR-Structural Engineering Research Center
Hammond, IN Chennai, Tamilnadu, India
Razaqpur, A. Ghani Saatci, Selcuk
McMaster University Izmir Institute of Technology
Hamilton, ON, Canada Izmir, Turkey
Regan, Paul Sabouni, Faisal
Trigram Architectural Consulting Group
London, UK Abu Dhabi, United Arab Emirates
Reiterman, Roy Sadeghi Pouya, Homayoon
Roy H. Reiterman, P.E., and Associates Coventry University
Troy, MI Coventry, UK
Ren, Xiaodan Saedi, Houman
Shanghai, China Tabiat Modares University
Richardson, James Tehran, Islamic Republic of Iran
University of Alabama Safan, Mohamed
Tuscaloosa, AL Menoufia University
Riding, Kyle Shebeen El-Koom, Menoufia, Egypt
Kansas State University Safi, Brahim
Manhattan, KS University of Boumerdes
Rinaldi, Zila Boumerdes, Algeria
University of Rome Tor Vergata Sagaseta, Juan
Rome, Italy University of Surrey
Rizk, Emad Guildford, Surrey, UK
Memorial University of Newfoundland Sagues, Alberto
St. John’s, NL, Canada University of South Florida
Rizwan, Syed Ali Tampa, FL
University of Engineering and Technology Sahamitmongkol, Raktipong
Lahore, Punjab, Pakistan CONTEC; Sirindhorn International Institute of Technology,

Roberts, Lawrence Thammasat University; MTEC
Grace Construction Products Pathumthani, Thailand
Cambridge, MA Sahmaran, Mustafa
Robery, Peter Gazi University
Halcrow Group Ltd. Ankara, Turkey
Solihull, West Midlands, UK Sahoo, Dipak
Cochin University of Science and Technology
Cochin, Kerala, India

478 ACI Structural Journal/March-April 2014


 REVIEWERS IN 2013
Saito, Shigehiko Sennour, Larbi
University of Yamanashi Consulting Engineers Group
Kofu, Japan San Antonio, TX
Sajedi, Fathollah Serna-Ros, Pedro
Universiti Malaya Universitat Politècnica de València
Kuala Lumpur, Selangor, Malaysia Valencia, Spain
Saka, Mehmet Serrano, Miguel
Middle East Technical University University of Oviedo
Ankara, Turkey Gijon, Spain
Sakai, Etsuo Shafigh, Payam
Tokyo Institute of Technology Kuala Lumpur, Malaysia
Ichikawa-shi, Japan Shafiq, Nasir
Saleem, Muhammad Universiti Teknologi Petronas
Florida International University Tronoh, Perak, Malaysia
Miami, FL Shah, Santosh
Salem, Hamed Dharmsinh Desai University
Cairo University Nadiad, Gujarat, India
Giza, Egypt Shah, Surendra
Sallam, Hossam El-Din Northwestern University
Zagazig University Evanston, IL
Zagazig, Sharkia, Egypt Shahnewaz, Md
Sánchez, Isidro University of British Columbia Okanagan
University of Alicante Kelowna, BC, Canada
Alicante, Alicante, Spain Shannag, M. Jamal
Sanchez, Leandro King Saud University
São Paulo, Brazil Riyadh, Saudi Arabia
Sant, Gaurav Shao, Yixin
University of California, Los Angeles McGill University
Los Angeles, CA Montreal, QC, Canada
Saqan, Elias Shariq, Mohd
American University in Dubai Civil Engineering
Dubai, United Arab Emirates Aligarh, Uttar Pradesh, India
Sarker, Prabir Sharma, Akanshu
Curtin University of Technology Bhabha Atomic Research Centre
Bentley, Western Australia, Australia Mumbai, Maharashtra, India
Sato, Ryoichi Shayan, Ahmad
Hiroshima University ARRB Group
Higashi-Hiroshima, Japan Vermont South, Victoria, Australia
Sato, Yuichi Shehata, Medhat
Kyoto University Ryerson University
Kyoto, Japan Toronto, ON, Canada
Scanlon, Andrew Sheikh, Shamim
The Pennsylvania State University University of Toronto
University Park, PA Toronto, ON, Canada
Schindler, Anton Sherman, Matthew
Auburn University Simpson Gumpertz & Heger
Auburn, AL Melrose, MA
Schwetz, Paulete Sherwood, Edward
Universidade Federal do Rio Grande do Sul Carleton University
Porto Alegre, Rio Grande do Sul, Brazil Ottawa, ON, Canada
Semaan, Hassnaa Shi, Xianming
Ottawa Hills, OH Bozeman, MT
Sener, Siddik Shi, Xudong
Gazi University Tsinghua University
Ankara, Turkey Beijing, China
Sengul, Ozkan Shield, Carol
Istanbul Technical University University of Minnesota
Istanbul, Turkey Minneapolis, MN

ACI Structural Journal/March-April 2014 479


REVIEWERS IN 2013
Shing, Pui-Shum Song, Kai
University of California, San Diego Building Materials
La Jolla, CA Dalian, China
Shivali, Ram Sossou, Gnida
Central Soil and Materials Research Station Kwame Nkrumah University of Science and Technology
New Delhi, India (KNUST)
Shrive, Nigel Kumasi, Ghana
University of Calgary Souza, Rafael
Calgary, AB, Canada Universidade Estadual de Maringá
Shukla, Abhilash Maringá, Paraná, Brazil
Jaypee University of Information Technology Söylev, Altug
Waknaghat, Himachal Pradesh, India Yeditepe University
Sideris, Kosmas Istanbul, Turkey
Democritus University of Thrace Sozen, Mete
Xanthi, Greece Purdue University
Sigrist, Viktor West Lafayette, IN
Hamburg University of Technology Spadea, Giuseppe
Hamburg, Germany University of Calabria
Sihotang, Fransiscus Arcavacata di Rende, Cosenza, Italy
National Taiwan University of Science and Technology
 Spinella, Nino
(NTUST) University of Messina
Taipei City, Taiwan Messina, Italy
Silfwerbrand, Johan Spyridis, Panagiotis
Swedish Cement and Concrete Research Institute Institute for Structural Engineering
Stockholm, Sweden Vienna, Austria
Singh, Harvinder Sreekala, R.
Guru Nanak Dev Engineering College Structural Engineering Research Centre
Ludhiana, Punjab, India Chennai, Tamilnadu, India
Sivey, Paul Stanton, John
Sivey Enterprises University of Washington
Hilliard, OH Seattle, WA
Skazlic, Marijan Stein, Boris
University of Zagreb Twining Laboratories
Zagreb, Croatia Long Beach, CA
Smith, Scott Steuck, Kyle
Southern Cross University University of Washington
Lismore, New South Wales, Australia Seattle, WA
Smyl, Danny Sudhahar, Sridevi
United States Marine Corps United Institute of Technology
Quantico, VA Coimbatore, Tamil Nadu, India
Sneed, Lesley Sujjavanich, Suvimol
Missouri University of Science and Technology Kasetsart University
Rolla, MO Bangkok, Thailand
So, Hyoung-Seok Sullivan, Patrick
Seonam University Sullivan and Associates
Namwon, Republic of Korea Rickmansworth, UK
Soejoso, Mia Sun, Shaoyun
Hiroshima University University of Illinois at Urbana-Champaign
Saijo, Japan Urbana, IL
Soltani, Amir Ta, Binh
Purdue University Calumet University of Civil Engineering
Hammond, IN Melbourne, Victoria, Australia
Soltanzadeh, Fatemeh Tabatabai, Habib
Engineering and Technology University of Wisconsin-Milwaukee
Aligarh, Uttar Pradesh, India Milwaukee, WI
Sonebi, Mohammed Tadayon, Mohammad Hosein
Queens University Belfast University of Tehran
Belfast, UK Tehran, Islamic Republic of Iran
Tadayon, Mohsen
Iranian Concrete Institute
Tehran, Islamic Republic of Iran

480 ACI Structural Journal/March-April 2014


 REVIEWERS IN 2013
Tadros, Maher Tavio
e.construct.USA, LLC Sepuluh Nopember Institute of Technology (ITS)
Omaha, NE Surabaya, East Java, Indonesia
Tae, Ghi ho Tegos, Ioannis
Leader Industrial Co. Salonica, Greece
Seoul, Republic of Korea Tehrani, Fariborz
Taghaddos, Hosein California State University, Fresno
PCL Industrial Management Inc. Fresno, CA
Edmonton, AB, Canada Thermou, Georgia
Tahmasebinia, Faham Aristotle University of Thessaloniki
University of Wollongong Thessaloniki, Greece
Wollongong, New South Wales, Australia Thiagarajan, Ganesh
Takahashi, Susumu University of Missouri-Kansas City
Nagoya Institute of Technology Kansas City, MO
Nagoya, Japan Thokchom, Suresh
Tan, Kefeng Manipur Institute of Technology
Southwest University of Science and Technology Imphal, India
Sichuan, China Thomas, Adam
Tan, Kiang Hwee Europoles GmbH & Co.
National University of Singapore Neumarkt, Germany
Singapore, Singapore Thompson, Phillip
Tanacan, Leyla Palm Desert, CA
Istanbul, Yesilkoy, Turkey Thorne, A.
Tanesi, Jussara Center of Engineering Materials and Structures
Federal Highway Administration-SaLUT Guilford, Surrey, UK
Vienna, VA Tian, Ying
Tang, Chao-Wei University of Nevada, Las Vegas
Cheng-Shiu University Las Vegas, NV
Niaosong District, Kaohsiung City, Taiwan Tiberti, Giuseppe
Tangtermsirikul, Somnuk University of Brescia
Sirindhorn International Institute of Technology Brescia, Italy
Patumthani, Thailand Tito, Jorge
Tank, Tejenadr University of Houston-Downtown
Pandit Deendayal Petroleum University Houston, TX
Gandhinagar, Gujarat, India Tixier, Raphael
Tankut, Tugrul Western Technologies Inc.
Middle East Technical University Phoenix, AZ
Ankara, Turkey Tjhin, Tjen
Tanner, Jennifer Buckland & Taylor Ltd
University of Wyoming North Vancouver, BC, Canada
Laramie, WY Tobolski, Matthew
Tantary, Manzoor San Diego, CA
Indian Institute of Technology Roorkee Tokgoz, Serkan
Roorkee, Uttarakhand, India Mersin University
Tapan, Mücip Mersin, Turkey
Yuzuncu Yil University Tolentino, Evandro
Van, Turkey Centro Federal de Educação Tecnológica de Minas Gerais
Tasdemir, Mehmet Timóteo, Minas Gerais, Brazil
Istanbul Technical University Topçu, İlker
Istanbul, Turkey Eskişehir Osmangazi University
Tassios, Theodosios Eskişehir, Turkey
Athens, Greece Torrenti, Jean-Michel
Tastani, S.P Chevilly Larue, France
Demokritus University of Thrace Torres-Acosta, Andres
Xanthi, Greece Universidad Marista de Querétaro
Tavares, Maria Querétaro, Mexico
State University of Rio de Janeiro (UERJ) Tosun, Kamile
Rio de Janeiro, Brazil Dokuz Eylül University
Izmir, Turkey

ACI Structural Journal/March-April 2014 481


REVIEWERS IN 2013
Toufigh, Vahab Vimonsatit, Vanissorn
University of Arizona Curtin University
Tucson, AZ Perth, Western Australia, Australia
Trautwein, Leandro Vintzileou, Elizabeth
Federal University of ABC National Technical University of Athens
São Paulo, Brazil Athens, Greece
Triantafillou, Thanasis Vitaliani, Renato
University of Patras University of Padua
Patras, Greece Padua, Italy
Tsonos, Alexander Viviani, Marco
Aristotle University of Thessaloniki Haute Ecole d’Ingénierie et de Gestion du Canton de Vaud
Thessaloniki, Greece (HEIG-VD)
Tsubaki, Tatsuya Yverdon les Bains, Switzerland
Yokohama National University Vogel, Thomas
Yokohama, Japan Institute of Structural Engineering
Tuchscherer, Robin Zurich, Switzerland
Northern Arizona University Vollum, Robert
Flagstaff, AZ Imperial College London
Tureyen, Ahmet London, UK
Wiss, Janney, Elstner Associates, Inc. Volz, Jeffery
Birmingham, MI The Pennsylvania State University
Turgut, Paki University Park, PA
Harran University Vosooghi, Ashkan
Sanliurfa, Turkey AECOM
Turk, A. Murat Sacramento, CA
Istanbul Kultur University Wagh, Prabhanjan
Istanbul, Turkey College of Engineering, Pune
Tutikian, Bernardo Satara, Maharashtra, India
Unisinos Waldron, Christopher
Porto Alegre, Rio Grande do Sul, Brazil University of Alabama at Birmingham
Unterweger, Andreas Birmingham, AL
Institute for Structural Engineering Wan, David
Vienna, Austria Old Castle Precast, Inc.
Uygunoglu, Tayfun South Bethlehem, NY
Afyon Kocatepe University Wang, Chang-Qing
Afyonkarahisar, Turkey Tongji University
Vakilly, Sedigheh Shanghai, China
Isfahan, Islamic Republic of Iran Wang, Chong
Varum, Humberto Brisbane, Queensland, Australia
University of Aveiro Wang, Huanzi
Aveiro, Portugal San Jose, CA
Vaz Rodrigues, Rui Wang, Junyan
École Polytechnique Fédérale de Lausanne (EPFL) National University of Singapore
Lausanne, VD, Switzerland Singapore
Vazquez-Herrero, Cristina Wang, Kejin
Universidade da Coruña Iowa State University
La Coruña, Spain Ames, IA
Veen, Cornelis Wehbe, Nadim
Delft University of Technology South Dakota State University
Delft, the Netherlands Brookings, SD
Velázquez Rodríguez, Sergio Wei, Ya
Universidad Panamericana University of Michigan
Zapopan, Jalisco, Mexico Ann Arbor, MI
Velu, Saraswathy Wei-Jian, Yi
Central Electro Chemical Research Institute Changsha, China
Karaikudi, Tamil Nadu, India Weiss, Charles
Vichit-Vadakan, Wilasa U.S. Army Corps of Engineers Engineer Research and

CTLGroup Development Center
Skokie, IL Vicksburg, MS

482 ACI Structural Journal/March-April 2014


 REVIEWERS IN 2013
Weiss, Jason Xin-hua, Cai
Purdue University Wuhan University
West Lafayette, IN Wuhan, HuBei, China
Wen, Ziyun Xu, Aimin
South China University of Technology ARRB Group
Guangzhou, Guangdong, China Melbourne, Victoria, Australia
Werner, Anne Xuan, D.X.
Southern Illinois University Edwardsville Delft University of Technology
Edwardsville, IL Delft, the Netherlands
Weyers, Richard Yahia, Ammar
Blacksburg, VA Université de Sherbrooke
Wheat, Harovel Sherbrooke, QC, Canada
University of Texas at Austin Yakoub, Haisam
Austin, TX Ottawa, ON, Canada
Wheeler, Andrew Yan, Libo
University Of Western Sydney University of Auckland
Sydney, New South Wales, Australia Auckland, New Zealand
Wilson, William Yang, KuoChen
Université de Sherbrooke National Kaohsiung First University of Science and Technology
Sherbrooke, QC, Canada Kaohsiung, Taiwan
Windisch, Andor Yang, Xinbao
Karlsfeld, Germany Olathe, KS
Won, Moon Yang, Yanan
Texas Tech University Pitt & Sherry
Lubbock, TX South Melbourne, Victoria, Australia
Wong, Hong Yang, Zhifu
Imperial College London Middle Tennessee State University
London, UK Murfreesboro, TN
Wong, Sook-Fun Yassein, Mohamed
Nanyang Technological University Doha, Qatar
Singapore Yatagan, Serkan
Wood, Richard Istanbul Technical University
University of California, San Diego Istanbul, Turkey
La Jolla, CA Yazıcı, Şemsi
Woyciechowski, Piotr Ege University
Warsaw University of Technology İzmir, Turkey
Warsaw, Poland Yehia, Sherif
Wu, Chenglin American University of Sharjah
Missouri University of Science and Technology Sharjah, United Arab Emirates
Rolla, MO Yen, Peter
Wu, Hui Bechtel National, Inc.
Beijing, China San Francisco, CA
Wu, Hwai-Chung Yerramala, Amarnath
Wayne State University Dundee University
Detroit, MI Dundee, Scotland, UK
Xia, Zuming Yigiter, Huseyin
Grand Prairie, TX Dokuz Eylül University
Xiang, Tianyu İzmir, Turkey
Chengdu, Sichuan, China Yildirim, Hakki
Xiangguo, Wu Istanbul, Turkey
Harbin Institute of Technology Yoon, Young-Soo
Harbin, Heilongjiang, China Korea University
Xiao, Feipeng Seoul, Republic of Korea
Clemson University Yost, Joseph
Clemson, SC Villanova University
Xiao, Yan Villanova, PA
Hunan University Youkhanna, Kanaan
Changsha, Hunan, China University of Dohuk
Xingyi, Zhu Duhok, Iraq
Hangzhou, China

ACI Structural Journal/March-April 2014 483


REVIEWERS IN 2013
Young-sun, Kim Zhang, Jun
Tokyo University of Science Tsinghua University
Noda-Shi, Chiba, Japan Beijing, China
Youssef, Maged Zhang, Peng
University of Western Ontario Karlsruhe Institute of Technology (KIT)
London, ON, Canada Karlsruhe, Germany
Yu, Baolin Zhang, Wei Ping
Michigan State University Tongji University
East Lansing, MI Shanghai, China
Yuan, Jiqiu Zhang, Xiaogang
PSI; Federal Highway Administration Turner-Fairbank Highway Shenzhen University
Research Center Shenzhen, Guangdong, China
McLean, VA Zhang, Xiaoxin
Yüksel, Isa Universidad de Castilla-La Mancha
Bursa Technical University Ciudad Real, Spain
Bursa, Turkey Zhang, Yamei
Zahedi, Farshad Southeast University
Babol Noshirvani University of Technology Nanjing, China
Babol, Mazandaran, Islamic Republic of Iran Zheng, Herbert
Zaki, Adel Gammon Construction Limited
SNC-Lavalin Hong Kong
Montreal, QC, Canada Zheng, Jianjun
Zanuy, Carlos Zhejiang University of Technology
Universidad Politécnica de Madrid Hangzhou, China
Madrid, Spain Zhou, Wei
Zatar, Wael Harbin Institute of Technology
West Virginia University Institute of Technology Harbin, China
Montgomery, WV Zhu, Han
Zerbino, Raul TianJin University
La Plata, Argentina TianJin, China
Zeris, Christos Ziehl, Paul
National Technical University of Athens University of South Carolina
Zografou, Greece Columbia, SC
Zhang, Jieying Zilch, Konrad
National Research Council Canada Technische Universität München
Ottawa, ON, Canada Munich, Germany

484 ACI Structural Journal/March-April 2014


2014 ACI Membership Application
American Concrete Institute • 38800 Country Club Drive • Farmington Hills, MI 48331 • USA
Phone: (248) 848-3800 • Fax: (248) 848-3801 • Web: www.concrete.org

Please print or type all information requested below:


 Mr.  Mrs.  Ms.  Dr. Gender:  Female  Male

FIRST NAME ______________________________________ MIDDLE INITIAL ___________ LAST NAME (SURNAME) ______________________________________________________

EMPLOYER _________________________________________________________________ CORPORATE TITLE __________________________________________________________

E-MAIL ADDRESS ____________________________________________________________ BIRTHDATE _________________________________________________________________

ADDRESS _______________________________________________________________________________________________________________________________________________

CITY _____________________________________________ STATE ___________________ ZIP __________________ COUNTRY ___________________________________________

TELEPHONE ________________________________________________________________ FAX ________________________________________________________________________

Categories of membership Included Subscriptions


Please select the desired category of membership and submit the appropriate dues described below.
Your ACI membership includes a hard copy and
 ORGANIZATIONAL – $1007 PLUS APPLICABLE SHIPPING FEES online subscription to Concrete International, and
Includes one set of the Manual of Concrete Practice (MCP), a wall plaque, and a subscription to
your choice of one of the following:
Concrete International, Concrete Repair Bulletin, and both ACI Materials and Structural Journals.
 ACI Materials Journal, hard copy
MANUAL OF CONCRETE PRACTICE: – Select a format and include applicable shipping fee:  ACI Materials Journal, online
 Seven Volume Set – U.S. add $30 shipping fee; Canada and Mexico add $130 shipping fee;  ACI Structural Journal, hard copy
and all others add $220 shipping fee.  ACI Structural Journal, online
 CD-ROM – no additional shipping costs  Concrete Repair Bulletin, hard copy
 Online Subscription – no additional shipping costs
 INDIVIDUAL – $226/year INDICATE SUBSCRIPTION PREFERENCE Please allow up to two months for delivery outside the U.S.
Individuals 28 years old or above residing worldwide.
 YOUNG PROFESSIONAL – $126/year INDICATE SUBSCRIPTION PREFERENCE To add subscriptions: please mark above and include an
Individuals under the age of 28 who do not qualify for student membership. additional $70 ($75 for individuals outside U.S. & Canada,
E-STUDENT – FREE Join at www.students.concrete.org. $41 for students outside U.S. & Canada, and $34 for U.S.
& Canada students) for hard copy subscriptions or $29 for
 STUDENT (U.S. AND CANADA) – $41/year INDICATE SUBSCRIPTION PREFERENCE
online subscriptions.
 STUDENT (Outside U.S. and Canada) – $81/year INDICATE SUBSCRIPTION PREFERENCE
Individuals under the age of 28 who are registered full-time students at an educational institution.
Full-time students age 28 and above may be granted Student Membership when the request is Optional Online Subscriptions
endorsed by the student’s advisor. Bundle your ACI membership with these optional
SUSTAINING MEMBERSHIP – ACI Sustaining Members receive all membership benefits of online subscriptions for additional ACI resources:
Organizational Members plus a free copy of every new ACI publication and increased corporate  Symposium Papers Subscription, $97
exposure, positioning them as a leader in the concrete industry, and much more. For complete details  Manual of Concrete Practice, $409
or to join, visit www.concrete.org/sustainingmembership or call (248) 848-3800.

Profile Information (Select all that apply.)


Markets Occupations
 Design  Management  Inspector  Researcher
 Materials  Consultant  Craftsman  Educator
 Production  Engineer  Sales & Marketing  Student
 Construction  Architect  Association  Other __________
 Testing  Contractor Employee
 Repair  Technical Specialist  Government
 Owner  Quality Control Employee

Fees Payment CHECK NUMBER

Membership dues  CHECK payable to American Concrete Institute


$
Additional subscriptions  CREDIT CARD  Visa  MasterCard  American Express
$
Organizational seven volume set
$ shipping fees
ACCOUNT NUMBER EXPIRATION DATE

$ For Journals, Canadian residents


add 5% GST (#126213149RT)
NAME (AS APPEARS ON CREDIT CARD)
$ TOTAL (U.S. Funds Only)
SIGNATURE
Code: CI

74 JANUARY 2014 Concrete international


ACI Structural Journal/March-April 2014485
NOTES:

486 ACI Structural Journal/March-April 2014


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

Then become a REVIEWER for the


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

Did you know that the database for MANUSCRIPT


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

QUESTIONS?
E-mail any questions to Journals.Manuscripts@concrete.org.
One Click... One Entire Journal
NOTES:

Introducing
a new feature on ACI’s website!

Now, with one click of the mouse, subscribers of the


ACI Structural and Materials Journals digital editions can
download an Adobe Acrobat PDF of an unabridged
issue at www.concrete.org/Publications/
ACIStructuralJournal.aspx and www.concrete.org/
Publications/ACIMaterialsJournal.aspx.
488 ACI Structural Journal/March-April 2014
ACI
STRUCTURAL
J O U R N A L
J O U R N

This journal and a companion periodical,


ACI Materials Journal, continue the publishing
tradition the Institute started in 1904.
Information published in ACI Materials Journal
includes: properties of materials used in
concrete; research on materials and concrete;
properties, use, and handling of concrete; and
related ACI standards and committee reports.

Das könnte Ihnen auch gefallen