Sei sulla pagina 1di 37

ACI 423.

5R-99

State-of-the-Art Report on
Partially Prestressed Concrete
Reported by Joint ACI-ASCE Committee 423

Ward N. Marianos, Jr.* Henry Cronin, Jr.


Chairman Secretary

Sarah L. Billington William L. Gamble H. Kent Preston


Kenneth B. Bondy Hans R. Ganz Denis C. Pu
Robert N. Bruce, Jr.* J. Weston Hall Julio A. Ramirez*
Dale Buckner Mohammad Iqbal Ken B. Rear
Ned H. Burns* Francis J. Jacques Dave Rogowsky
Gregory P. Chacos* Daniel P. Jenny Bruce W. Russell
Jack Christiansen Paul Johal David H. Sanders
Todd Christopherson Susan N. Lane Thomas Schaeffer*
Steven R. Close * Morris Schupack
Les Martin
Thomas E. Cousins Alan H. Mattock* Kenneth W. Shushkewich
* Gerard J. McGuire Khaled S. Soubra
Charles W. Dolan
Apostolos Fafitis Mark Moore* Richard W. Stone
Mark W. Fantozzi Antoine E. Naaman* Patrick Sullivan
Martin J. Fradua Kenneth Napior Luc R. Taerwe
Catherine W. French* Thomas E. Nehil H. Carl Walker
Clifford Freyermuth Mrutyunjaya Pani Jim J. Zhao
Paul Zia*

*Subcommittee preparing report (Michael Barker contributed to writing Chapters 4 and 5 of this report).

Partially prestressed concrete construction uses prestressed, or a combina- partially prestressed. This report is presented as an overview of the current
tion of prestressed and nonprestressed, reinforcement. Partially prestressed state of the art for partial prestressing of concrete structures. Research
concrete falls between the limiting cases of conventionally reinforced con- findings and design applications are presented. Specific topics discussed
crete and fully prestressed concrete, which allows no flexural tension under include the history of partial prestressing, behavior of partially prestressed
service loads. When flexural tensile stresses and cracking are allowed concrete members under static loads, time-dependent effects, fatigue, and
under service loads, the prestressed members have historically been called the effects of cyclic loadings.

ACI Committee Reports, Guides, Standard Practices, and Keywords: bridges; buildings; concrete construction; corrosion; cracking;
Commentaries are intended for guidance in planning, crack widths; cyclic loading; deflections; earthquake-resistant structures;
designing, executing, and inspecting construction. This fatigue; partially prestressed concrete; post-tensioning; prestressing; pre-
document is intended for the use of individuals who stress losses; shear; stresses; structural analysis; structural design; time-
are competent to evaluate the significance and dependent effects; torsion.
limitations of its content and recommendations and
who will accept responsibility for the application of the CONTENTS
material it contains. The American Concrete Institute Chapter 1—Introduction, p. 423.5R-2
disclaims any and all responsibility for the stated 1.1—Historical perspective
principles. The Institute shall not be liable for any loss or 1.2—Definition
damage arising therefrom. 1.3—Design philosophy of partial prestressing
Reference to this document shall not be made in
contract documents. If items found in this document are ACI 423.5R-99 became effective December 3, 1999.
Copyright  2000, American Concrete Institute.
desired by the Architect/Engineer to be a part of the All rights reserved including rights of reproduction and use in any form or by any
contract documents, they shall be restated in mandatory means, including the making of copies by any photo process, or by electronic or
mechanical device, printed, written, or oral, or recording for sound or visual reproduc-
language for incorporation by the Architect/Engineer. tion or for use in any knowledge or retrieval system or device, unless permission in
writing is obtained from the copyright proprietors.

423.5R-1
423.5R-2 ACI COMMITTEE REPORT

1.4—Advantages and disadvantages of partial (Freyssinet 1933). Freyssinet recognized that as the load on
prestressing a prestressed member is increased, flexural cracks would
1.5—Partial prestressing and reinforcement indexes appear in the tensile zones at a certain load level, which he
1.6—Report objective referred to as the transformation load. Even though the
cracks would close as the load was reduced and the structure
Chapter 2—Partially prestressed members under would recover its original appearance, Freyssinet advocated
static loading, p. 423.5R-5 avoiding cracks under service load so that the concrete
2.1—Behavior would behave as a homogeneous material.
2.2—Methods of analysis
A different design approach, however, was proposed by
2.3—Cracking
von Emperger (1939) and Abeles (1940). They suggested
2.4—Deflections
using a small amount of tensioned high-strength steel to
2.5—Shear and torsion
control deflection and crack width while permitting higher
Chapter 3—Time-dependent behavior, p. 423.5R-12 working stresses in the main reinforcement of reinforced
3.1—Prestress losses concrete. Most of the early work in support of this design
3.2—Cracking concept was done by Abeles (1945) in England. Based on his
3.3—Deflections studies, Abeles determined that eliminating the tensile stress
3.4—Corrosion and possible cracking in the concrete is unnecessary in many
designs. Abeles also realized that prestress can be applied to
Chapter 4—Effects of repeated loading (fatigue), counteract only part of the service load so that tensile stress,
p. 423.5R-15 or even hairline cracks, occur in the concrete under full
4.1—Background service load. Abeles did specify that under dead load only,
4.2—Material fatigue strength no flexural tension stress should be allowed at any member
4.3—Fatigue in partially prestressed beams face where large flexural tensile stresses occurred under
4.4—Prediction of fatigue strength maximum load, so as to ensure closure of any cracks that
4.5—Serviceability aspects may have occurred at maximum load. Additional bonded
4.6—Summary of serviceability and well-distributed nonprestressed reinforcement could be
used to help control cracking and provide the required
Chapter 5—Effects of load reversals, p. 423.5R-20 strength. Abeles termed this design approach as “partially
5.1—Introduction
prestressed concrete.” Therefore, the design approach advo-
5.2—Design philosophy for seismic loadings
cated by Freyssinet was then termed as “fully prestressed
5.3—Ductility
concrete.” In actual practice, nearly all prestressed concrete
5.4—Energy dissipation
components designed today would be “partially prestressed”
5.5—Dynamic analyses
as viewed by Freyssinet and Abeles.
5.6—Connections
5.7—Summary Interest in partial prestressing continued in Great Britain in
the 1950s and early 1960s. Many structures were designed
Chapter 6—Applications, p. 423.5R-28 by Abeles based on the principle of partial prestressing, and
6.1—Early applications examinations of most of these structures around 1970
6.2—Pretensioned concrete components revealed no evidence of distress or structural deterioration,
6.3—Post-tensioned building construction as discussed in the technical report on Partial Prestressing
6.4—Bridges published by the Concrete Society (1983). Partially
6.5—Other applications prestressed concrete design was recognized in the First
Report on Prestressed Concrete published by the Institution
Chapter 7—References, p. 423.5R-30 of Structural Engineers (1951). Provisions for partial
7.1—Referenced standards and reports prestressing were also included in the British Standard Code
7.2—Cited references of Practice for Prestressed Concrete (CP 115) in 1959. In that
code, a permissible tensile stress in concrete as high as 750
Appendix—Notations, p. 423.5R-36 psi (5.2 MPa) was accepted when the maximum working
load was exceptionally high in comparison with the load
CHAPTER 1—INTRODUCTION normally carried by the structure. Presently, the British Code
1.1—Historical perspective
Application of prestressing to concrete members imparts a (BS 8110) as well as the Model Code for Concrete Structures
compressive force of an appropriate magnitude at a suitable (1978), published by CEB-FIP, defines three classes of
location to counteract the service-load effects and modifies prestressed concrete structures:
the structural behavior of the members. Although the con- Class 1—Structures in which no tensile stress is permitted
cept of prestressed concrete was introduced almost concur- in the concrete under full service load;
rently in the U.S. and in Germany before the turn of the 20th Class 2—Structures in which a limited tensile stress is per-
century (Lin and Burns 1981), its principle was not fully mitted in the concrete under full service load, but there is no
established until Freyssinet published his classical study visible cracking; and
PARTIALLY PRESTRESSED CONCRETE 423.5R-3

Class 3—Structures in which cracks of limited width illustrated in that guide. The second edition (1978) mentioned
(0.2 mm [0.008 in.]) are permitted under full service load. the term “partial prestressing,” and by the third edition (1985),
Calculations for Class 3 structures would be based on the design examples of members with combined prestressed and
hypothetical tensile stress in the concrete assuming an nonprestressed reinforcement were included. Presently, ACI
uncracked section. The allowable values of the hypothetical 318 permits a tensile stress limit of 12√f c′ psi with
tensile stress vary with the amount, type, and distribution of requirements for minimum cover and a deflection check.
the prestressed and nonprestressed reinforcement. Section 18.4.3 of ACI 318 permits the limit to be exceeded on
Elsewhere in Europe, interest in partial prestressing also the basis of analysis or test results. Bridge design guidelines or
developed in the 1950s and 1960s. In the mid-1950s, many recommendations, however, did not follow the development
prestressed concrete structures in Denmark, especially until the publication of the Final Draft LRFD Specifications
bridges, were designed using the partial prestressing concept. for Highway Bridges Design and Commentary (1993), even
Their performance was reported as satisfactory after 25 years though most bridge engineers had been allowing tension in
of service (Rostam and Pedersen 1980). In 1958, the first their designs for many years.
partially prestressed concrete bridge in Switzerland The concept of partial prestressing was developed half a
(Weinland Bridge) was completed near Zurich. Provisions century ago. Over the years, partial prestressing has been
for partial prestressing were introduced in SIA Standard 162, accepted by engineers to the extent that it is now the normal
issued by the Swiss Society of Engineers and Architects way to design prestressed concrete structures. Bennett’s
(1968), and since 1960, more than 3000 bridges have been work (1984) provides a valuable historical summary of the
designed according to this concept with highly satisfactory development of partially prestressed concrete.
results (Birkenmaier 1984). Unlike the British Code and
CEP-FIP Model Code, the limit of partial prestressing in the 1.2—Definition
Swiss Code was not defined by the hypothetical tensile Despite a long history of recognition of the concept of
stress. Instead, it was defined by the tensile stress in the partial prestressing, both in the U.S. and abroad, there has
prestressed and nonprestressed reinforcement, and been a lack of a uniform and explicit definition of the term,
calculated using the cracked section. Under full service “partial prestressing.” For example, Lin and Burns (1981)
load, the allowable stress in the nonprestressed state: “When a member is designed so that under the working
reinforcement was 22,000 psi (150 MPa), and in railroad load there are no tensile stresses in it, then the concrete is
bridges, the stress increase in the prestressed reinforcement said to be fully prestressed. If some tensile stresses will be
was not to exceed 1/20 of the tensile strength. This value was produced in the member under working load, then it is
taken as 1/10 of the tensile strength in other structures. It was termed partially prestressed.” On the other hand, Naaman
required, however, that the concrete be in compression when (1982a) states: “Partial prestressing generally implies a com-
the structure supported only permanent load. bination of prestressed and nonprestressed reinforcement,
In the U.S., the design of prestressed concrete in the early both contributing to the resistance of the member. The aim is
1950s was largely based on the Criteria for Prestressed Con- to allow tension and cracking under full service loads while
crete Bridges (1954) published by the Bureau of Public ensuring adequate strength.” According to Nilson (1987),
Roads, which did not permit tensile stress and cracking in “Early designers of prestressed concrete focused on the com-
concrete under service loads. The ACI-ASCE Joint Commit- plete elimination of tensile stresses in members at normal
tee 323 report (1958), however, recognized that “complete service load. This is defined as full prestressing. As experi-
freedom from cracking may or may not be necessary at any ence has been gained with prestressed concrete construction,
particular load stage.” For bridge members, tensile stress it has become evident that a solution intermediate between
was not allowed in concrete subjected to full service load. full prestressed concrete and ordinary reinforced concrete
For building members not exposed to weather or corrosive offers many advantages. Such an intermediate solution, in
atmosphere, a flexural tension stress limit of 6√f c′ psi* was which a controlled amount of concrete tension is permitted
specified with the provision that the limit may be exceeded at full service, is termed partial prestressing.”
if “it is shown by tests that the structure will behave properly A unified definition of the term “partial prestressing”
under service load conditions and meet any necessary should be based on the behavior of the prestressed member
requirements for cracking load or temporary overload.” under a prescribed loading. Therefore, this report defines
Thus, partial prestressing was permitted in that first defini- partial prestressing as: “An approach in design and construc-
tive design guide for prestressed concrete, and designers tion in which prestressed reinforcement or a combination of
were quick to embrace the idea. When the balanced load prestressed and non-prestressed reinforcement is used such
design concept was published by Lin (1963), it provided a that tension and cracking in concrete due to flexure are
convenient design tool and encouraged the practical applica- allowed under service dead and live loads, while serviceabil-
tion of partial prestressing. ity and strength requirements are satisfied.”
In 1971, the first edition of the PCI Design Handbook was For the purposes of this report, fully prestressed concrete
published. Design procedures allowing tension stresses are is defined as concrete with prestressed reinforcement and no
flexural tension allowed in the concrete under service loads.
* In this report, when formulas or stress values are taken directly from U.S. codes Conventionally reinforced concrete is defined as concrete
and recommendations, they are left in U.S. customary units. with no prestressed reinforcement and generally, there is
423.5R-4 ACI COMMITTEE REPORT

flexural tension in concrete under service loads. Partially multispan bridges, camber control is important in improving
prestressed concrete falls between these two limiting cases. riding comfort as a vehicle passes from one span to the next.
Serviceability requirements include criteria for crack widths, The relatively large mild steel bars used in partially pre-
deformation, long-term effects (such as creep and shrink- stressed members result in a transformed section that can be
age), and fatigue. significantly stiffer than a comparable section that relies
By the previous definition, virtually all prestressed con- solely on prestressing strand, thus reducing both camber and
crete that uses unbonded tendons is “partially prestressed,” deflection.
as codes require that a certain amount of bonded reinforce- Nonprestressed reinforcement used in partially prestressed
ment be provided to meet strength requirements. Most pre- members will enhance the strength and also control crack
tensioned members used in routine applications such as formation and crack width. Under ultimate load, a partially
building decks and frames, and bridges spanning to approx- prestressed member usually demonstrates greater ductility
imately 100 ft (30 m) will allow flexural tension under full than a fully prestressed member. Therefore, it will be able to
service load. The addition of nonprestressed reinforcement is absorb more energy under extreme dynamic loading such as
used only in special situations, such as unusually long spans an earthquake or explosion.
or high service loads, or where camber and deflection control Because mild steel does not lose strength as rapidly as pre-
is particularly important. stressing strands at elevated temperature, it is sometimes
added to prestressed members to improve their fire-resis-
1.3—Design philosophy of partial prestressing tance rating. See Chapter 9 of the PCI Design Handbook
The basic design philosophy for partial prestressing is not (1992) and Design for Fire Resistance of Precast Pre-
different from that of conventionally reinforced concrete or stressed Concrete (1989) for more information.
fully prestressed concrete. The primary objective is to pro-
Partial prestressing is not without some disadvantages.
vide adequate strength and ductility under factored load and
Under repeated loading, the fatigue life of a partially pre-
to achieve satisfactory serviceability under full service load.
stressed member can be a concern. In addition, durability is
By permitting flexural tension and cracking in concrete, a potential problem for partially prestressed members
the designer has more latitude in deciding the amount of pre- because they can be cracked under full service load. Recent
stressing required to achieve the most desirable structural studies (Harajli and Naaman 1985a; Naaman 1989; and Naa-
performance under a particular loading condition. Therefore, man and Founas 1991), however, have shown that fatigue
partial prestressing can be viewed as a means of providing strength depends on the range of stress variation of the strand
adequate control of deformation and cracking of a pre- (refer to Chapter 4) and that durability is related more to cov-
stressed member. If the amount of prestressed reinforcement er and spacing of reinforcement than to crack width, so these
used to provide such control is insufficient to develop the concerns can be addressed with proper design and detailing
required strength, then additional nonprestressed reinforce- of the reinforcement (Beeby 1978 and 1979).
ment is used.
In the production of precast, pretensioned concrete mem-
1.5—Partial prestressing and reinforcement
bers, serviceability can be improved by placing additional indexes
strands, as this is more economical than placing reinforcing Several indexes have been proposed to describe the extent
bars. When this technique is used, the level of initial pre- of prestressing in a structural member. These indexes are
stress in some or all of the strands is lowered. This is also a useful in comparing relative performances of members made
useful technique to keep transfer stresses below the maxi- with the same materials, but caution should be exercised in
mum values prescribed by codes. At least for purposes of using them to determine absolute values of such things as
shear design, the ACI Building Code treats any member with deformation and crack width. Two of the most common indi-
effective prestress force not less than 40% of the tensile ces are the degree of prestress λ, and the partial prestressing
strength of the flexural reinforcement as prestressed concrete. ratio (PPR). These indexes are defined as

1.4—Advantages and disadvantages of partial


prestressing M dec
λ = ---------------------
- (1-1)
In the design of most building elements, the specified live MD + ML
load often exceeds the normally applied load. This is to
account for exceptional loading such as those due to impact, where
extreme temperature and volume changes, or a peak live
Mdec = decompression moment (the moment that produces
load substantially higher than the normal live loads. By
zero concrete stress at the extreme fiber of a section,
using partial prestressing, and by allowing higher flexural
nearest to the centroid of the prestressing force,
tension for loading conditions rarely imposed, a more eco-
when added to the action of the effective prestress
nomical design is achieved with smaller sections and less
alone);
reinforcement.
Where uniformity of camber among different members of MD = dead-load moment; and
a structure is important, partial prestressing will enable the ML = live-load moment
designer to exercise more control of camber differentials. In and
PARTIALLY PRESTRESSED CONCRETE 423.5R-5

M np CHAPTER 2—PARTIALLY PRESTRESSED


PPR = --------
- (1-2) MEMBERS UNDER STATIC LOADING
Mn 2.1—Behavior
There are a number of investigations on the behavior of
where partially prestressed concrete beams under static loading
Mnp = nominal moment capacity provided by prestressed (Abeles 1968; Burns 1964; Cohn and Bartlett 1982; Harajli
reinforcement; and 1985; Harajli and Naaman 1985a; Shaikh and Branson 1970;
Mn = total nominal moment capacity. Thompson and Park 1980a; and Watcharaumnuay 1984). The
In the previous expressions, all moments are computed at following results were observed for beams having the same
critical sections. This report will generally use the PPR to ultimate resistance in flexure but reinforced with various
describe the extent of prestressing in flexural members. The combinations of prestressed and nonprestressed reinforcement:
tests, studies, and examples described in this report usually • Partially prestressed beams show larger ultimate deflec-
concern members with PPR < 1, and the members are pre- tions, higher ductility, and higher energy absorption than
tensioned unless otherwise noted. fully prestressed beams;
Characterizing the total amount of flexural reinforcement • Partially prestressed beams tend to crack at lower load
in a member is also important. This will be done with the levels than fully prestressed beams. Average crack spac-
reinforcement index ω ing and crack widths are smaller. The stiffness of par-
tially prestressed beams after cracking is larger;
• For a given reinforcement index ω, the moment-curva-
f f ps f
ω = ρ ----y- + ρ p ----- – ρ′ ----y- (1-3) ture relationship is almost independent of the ratio of the
f c′ f c′ f c′ tensile reinforcement areas (prestressed versus nonpre-
stressed);
where • Changing the effective prestress in the prestressing ten-
dons does not lead to any significant change in the ulti-
mate resistance and curvature of flexural members; and
A
ρ = -----s- • A decrease in effective prestress leads to an increase in
bd yield curvature and a decrease in curvature ductility.

A′ 2.2—Methods of analysis
ρ′ = ------s- Several methods can be followed to analyze partially
bd
prestressed concrete members subjected to bending. In terms
of assumptions, purpose, and underlying principles, they are
A ps identical to those used for reinforced and prestressed
ρ p = -------
-
bd p concrete (Nilson 1976, Naaman and Siriaksorn 1979,
and Siriaksorn and Naaman 1979, Al-Zaid and Naaman 1986,
Aps = area of prestressed reinforcement in tension zone, and Tadros 1982).
in.2 (mm2); 2.2.1 Linear elastic analysis—In the elastic range of
As = area of nonprestressed tension reinforcement, in.2 behavior, the analysis must accommodate either a cracked or
(mm2); an uncracked section subjected to bending, with or without
A′s = area of nonprestressed compression reinforce- prestress in the steel. The usual assumptions of plane strain
ment, in.2 (mm2); distribution across the section, linear stress-strain relations,
b = width of compression face of member, in. (mm); and perfect bond between steel and concrete remain applica-
d = distance from extreme compression fiber to cen- ble. Linear elastic analysis under service loads assuming an
troid of nonprestressed tension reinforcement, in. uncracked section is used for prestressed concrete. In the
(mm); U.S., the design of reinforced concrete is predominantly
dp = distance from extreme compression fiber to cen- based on strength requirement, but a linear elastic analysis
troid of prestressed reinforcement, in. (mm); under service loads is also necessary to check serviceability
f ′c = specified compressive strength of concrete, psi limitations such as crack widths, deflections, and fatigue.
(MPa); Prestressed concrete beams can act as cracked or
fps = stress in prestressed reinforcement at nominal uncracked sections, depending on the level of loading. In
strength, psi (MPa); and contrast to reinforced concrete, the centroidal axis of the
fy = yield strength of nonprestressed reinforcement, psi cracked section does not coincide with the neutral axis point
(MPa). of zero stress (Fig. 2.1). Moreover, the point of zero stress does
not remain fixed, but moves with a change in applied load.
1.6—Report objective When the effective prestress tends toward zero, the point of
The objective of this report is to summarize the state of the zero stress and the centroidal axis tend to coincide. Generalized
art of the current knowledge as well as recent developments equations have been developed to determine the zero stress
in partial prestressing so that engineers who are not experi- point based on satisfying equilibrium, strain compatibility, and
enced in prestressed concrete design will have a better stress-strain relations (Nilson 1976; Naaman and Siriaksorn
understanding of the concept. 1979; Siriaksorn and Naaman 1979; and Al-Zaid and Naaman
423.5R-6 ACI COMMITTEE REPORT

provide unified treatment for cracked reinforced, prestressed,


and partially prestressed sections.
2.2.2 Strength analysis—At ultimate or nominal moment
resistance, the assumptions related to the stress and strain
distributions in the concrete, such as the compression block
in ACI 318, or the stress and strain in the steel (such as yield-
ing of the reinforcing steel) are identical for reinforced, pre-
stressed, and partially prestressed concrete (Fig. 2.2). The
corresponding analysis is the same and leads to the nominal
moment resistance of the section. Numerous investigations
have shown close correlation between the predicted (based
on ACI 318) and experimental values of nominal moments.
The ACI 318 analysis, however, resulted in conservative
predictions of section curvatures at ultimate load, leading to
erroneous estimates of deformations and deflections (Wang
et al. 1978, Naaman et al. 1986). To improve the prediction
of nominal moment and curvature, either a nonlinear or a
simplified nonlinear analysis may be followed.
Simplified nonlinear analysis—In the simplified nonlinear
analysis procedure (also called pseudo-nonlinear analysis),
Fig. 2.1—Assumed stress or strain distribution in linear the actual stress-strain curve of the steel reinforcement is
elastic analysis of cracked and uncracked sections (Naaman considered while the concrete is represented by the ACI 318
1985).
compression block. A solution can be obtained by solving
two nonlinear equations with two unknowns, namely the
stress and the strain in the prestressing steel at nominal
moment resistance (Naaman 1977, Naaman 1983b).
Nonlinear analysis—The best accuracy in determining
nominal moments and corresponding curvatures is achieved
through a nonlinear analysis procedure (Cohn and Bartlett
1982, Naaman et al. 1986, Harajli and Naaman 1985b,
Moustafa 1986). Nonlinear analysis requires as input an
accurate analytical representation of the actual stress-strain
curves of the component materials (concrete, reinforcing
steel, and prestressing steel). Typical examples can be found
in two references (Naaman et al. 1986, Moustafa 1986).

2.3—Cracking
Partially prestressed concrete permits cracking under ser-
vice loads as a design assumption. To satisfy serviceability
requirements, the maximum crack width should be equal to, or
smaller than, the code-recommended limits on crack width.
The maximum allowable crack widths recommended by
ACI Committee 224 (1980) for reinforced concrete members
can be used, preferably with a reduction factor for pre-
stressed and partially prestressed concrete members. To
select the reduction factor, consideration should be given to
the small diameter of the reinforcing elements (bars or
strands), the cover, and the exposure conditions.
Only a few formulas are used in the U.S. practice to predict
crack widths in concrete flexural members. Because the fac-
Fig. 2.2—Assumed strain distribution and forces in: (a)
nonlinear analysis; (b) approximate nonlinear analysis; tors influencing crack widths are the same for reinforced and
and (c) ultimate strength analysis by ACI Code (Naaman partially prestressed concrete members, existing formulas
1985). for reinforced concrete can be adapted to partially pre-
stressed concrete. Five formulas (ACI 224 1980; Gergely
1986). They usually are third-order equations with respect to and Lutz 1968; Nawy and Potyondy 1971; Nawy and Huang
member depth. Although they can be solved iteratively, charts, 1977; Nawy and Chiang 1980; Martino and Nilson 1979; and
tables, and computer programs have been developed for their Meier and Gergely 1981) applicable to partially prestressed
solution (Tadros 1982, Moustafa 1977). These equations beams are summarized in Table 2.1 (Naaman 1985). The vari-
PARTIALLY PRESTRESSED CONCRETE 423.5R-7

Table 2.1—Crack width prediction equations applicable to partially prestressed beams (Naaman 1985)
Source Equation* with U.S. system, (in., ksi) Equation* with SI system, (mm., N/mm2)

W max = 7.6 × 10 βf s 3 d c A b Multiply expression by 0.1451


–5

fs = tensile stress in reinforcing steel


Gergely and Lutz (1968)
ACI Code (1971, 1977, and 1983) dc = concrete cover to center of closest bar
ACI Committee 224 (1980) layer Note: ACI Committee 224 recommends
Ab = concrete tensile area per bar multiplication factor of 1.5 when strands,
rather than deformed bars, are used nearest
β = ratio of distances from tension face and to beam tensile face.
steel centroid to neutral axis
W max = 1.44 × 10 ( f s – 8.3 ) = 5.31 × 10 ( f s – 57.2 )
–4 –4
Nawy and Potyondy (1971)

A
W max = α × 10 β -------t- ∆f ps
–5
ΣO
At = area of concrete tensile zone
ΣO = sum of perimeters of bonded
Nawy and Huang (1977) reinforcing elements
Nawy and Chiang (1980) ∆fps = net stress change in prestressing steel Multiply expression by 0.1451
after decompression
5.85 if pretensioning
α= 
 6.51 if post-tensioning

W max = 14 × 10 d′c f s + 0.0031 = 2 × 10 d′c f s + 0.08


–5 –5

Martino and Nilson (1979)


d′c= concrete clear cover

 W max = C 1 ε ct d c (1) (1) Same equation



 W max = C 2 ε ct d c 3 A b (2) (2) Multiply by 220
Meier and Gergely (1981) C1, C2 = bond coefficients
For reinforcing bars: C1 = 12; C2 = 8.4
For strands: C1 = 16; C2 = 12
εct = nominal concrete tensile strain at
tensile face
*
In the formulas shown, fs can be replaced by ∆fps when applied to partially prestressed concrete.

able tensile stress in the reinforcing steel fs should be replaced ng (1977), and Meier and Gergely (1981). Although none of
by the stress change in the prestressing steel after decompres- the three equations gave sufficiently good correlation with
sion ∆fps. The ACI 318 formula initially developed by experimental data for all conditions, the following observa-
Gergely and Lutz (1968) for reinforced concrete could be tions were made (Fig. 2.3):
used as a first approximation for partially prestressed con- • The Gergely and Lutz equation gave a lower prediction
crete. Meier and Gergely (1981), however, suggested a mod- in all cases (Fig. 2.3(a));
ified form (shown in Table 2.1) for the case of prestressed • The Meier and Gergely equation gave the worst corre-
concrete. This alternate formula uses the nominal strain at lation (Fig. 2.3(c)); and
the tensile face of the concrete (instead of the stress in the • The Nawy and Huang equation gave a higher prediction
steel), and the cover to the center of the steel dc. Both the in most cases (Fig. 2.3(b)).
stress in the steel and the clear concrete cover are found to be Although more experimental data are needed to improve
the controlling variables in the regression equation derived the accuracy of crack-width prediction equations available in
by Martino and Nilson (1979). The two prediction equations U.S. practice, there is sufficient information to judge if the
proposed by Nawy and Huang (1977) and Nawy and Chiang serviceability, with respect to cracking or crack width under
(1980) contain most of the important parameters found in the short-term loading, is satisfactory for a partially prestressed
cracking behavior of concrete members except the concrete member. The effects of long-term loading and repetitive
loading (fatigue) on the crack widths of partially prestressed
cover, which is accounted for indirectly. Moreover, they are
members need to be further clarified. A research investiga-
based on actual experimental results on prestressed and par-
tion provided an analytical basis to deal with the problem
tially prestressed beams.
(Harajli and Naaman 1989); however, the proposed method-
As pointed out by Siriaksorn and Naaman (1979), large ology is not amenable to a simple prediction equation that
differences can be observed in predicted crack widths can be easily implemented for design.
depending on the prediction formula used. Harajli and Naa-
man (1989) compared predicted crack widths with observed 2.4—Deflections
crack widths from tests on twelve partially prestressed con- Fully prestressed concrete members are assumed to be
crete beams. They considered the three prediction equations uncracked and linearly elastic under service loads. Instantaneous
recommended by Gergely and Lutz (1968), Nawy and Hua- short-term deflections are determined using general
423.5R-8 ACI COMMITTEE REPORT

The widely accepted concept of the effective moment of


inertia Ieff , initially introduced by Branson (1977) for rein-
forced concrete, has been examined by several researchers
and modified accordingly to compute the deflection in
cracked prestressed and partially prestressed members. The
modified effective moment of inertia is defined (Naaman
1982a) as

3
I eff = I cr + ( ( M cr – M dec ) ⁄ ( M a – M dec ) ) (2-1)
( I g – I cr ) ≤ I g
where
Ig = gross moment of inertia, in.4 (mm4);
Icr = moment of inertia of cracked section, in.4 (mm4);
Mcr = cracking moment, in.-k (mm-N);
Mdec = decompression moment, in.-k (m-N); and
Ma = applied moment, in.-k (m-N).
Although there is general agreement for the use of the pre-
vious expression, substantial divergence of opinion exists as
to the computation of Icr and Mdec. The computation differ-
ence is whether the moment of inertia of the cracked section
should be computed with respect to the neutral axis of bend-
ing or with respect to the zero-stress point, and whether the
decompression moment should lead to decompression at the
extreme concrete fiber or whether it should lead to a state of
zero curvature in the section. The discussion of Tadros’
paper (1982) by several experts in the field is quite informa-
tive on these issues. A systematic comparison between the
various approaches, combined with results from experimen-
tal tests, is given in work by Watcharaumnuay (1984), who
observed that the use of Icr with respect to the neutral axis of
bending is preferable, while the use of Mdec as that causing
decompression at the extreme concrete fiber, is easier and
leads to results similar to those obtained using the zero cur-
vature moment.

2.5—Shear and torsion


2.5.1 General—Nonprestressed and fully prestressed
concrete (tensile stress in the concrete under full service load
Fig. 2.3—Comparison of observed and theoretically predicted is zero) are the two limiting cases of steel-reinforced con-
crack widths (Naaman 1985). crete systems. Partially prestressed concrete represents a
continuous transition between the two limit cases. A unified
principles of mechanics. To compute short-term deflections, approach in design to combined actions including partial pre-
customary U.S. practice is to use the gross moment of inertia stressing would offer designers a sound basis to make the
Ig for pretensioned members, or the net moment of inertia In appropriate choice between the two limits (Thurlimann 1971).
for members with unbonded tendons, and the modulus of The equivalent load concept provides a simple and
elasticity of concrete at time of loading or transfer Eci. efficient design of prestressed concrete structures under
Several approaches proposed by various researchers to combined actions (Nilson 1987). For example, this approach
compute short-term and long-term deflections in prestressed allows the designer to calculate the shear component of the
or partially prestressed uncracked members are summarized prestress anywhere in the beam, simply by drawing the shear
in Table 2.2 (Branson and Kripanarayanan 1971; Branson diagram due to the equivalent load resulting from a change
1974; Branson 1977; Naaman 1982a; Naaman 1983a; in the vertical alignment of the tendon (Fig. 2.4). That
Branson and Trost 1982a; Branson and Trost 1982b; Martin equivalent load, together with the prestressing forces acting
1977; Tadros et al. 1975; Tadros et al. 1977; Dilger 1982; at the ends of the member through the tendon anchorage,
and Moustafa 1986). Although no systematic evaluation or may be looked upon as just another system of external forces
comparison of these different approaches has been acting on the member. This procedure can be used for both
undertaken, for common cases they lead to results of the statically determinate and indeterminate structures, and it
same order. accounts for the effects of secondary reactions due to
PARTIALLY PRESTRESSED CONCRETE 423.5R-9

Table 2.2—Deflection prediction equations for prestressed and partially prestressed beams (from Naaman
1985)
Short-term instantaneous Long-term or additional long-term
Source deflection deflection Remarks
Long-term deflection obtained by
∆t is obtained from elastic integrating curvatures with due • Uncracked section; and
ACI 435 (1963) analysis using Ft , Ect , and Ig . account for creep effects and prestress • No provisions for As and A′s.
losses with time.
∆t shall be obtained from ∆add shall be computed, taking into • No provisions for partial
ACI Code Section 9.5 account stresses under sustained load,
elastic analysis using Ig for prestressing (cracking, As
(1971, 1977, and 1983) including effects of creep, shrinkage,
uncracked sections. and relaxation. and A′s ).

1+η
∆ add = η – 1 +  ------------- k r C cu
 2 
( ∆ t ) F + k r C CU ( ∆ t ) G + K CA k r C CU ( ∆ t ) SD • Uncracked section;
t
∆t is obtained from elastic • k r is applicable only when
Branson et al. (1971, where
1974, and 1977) CCU = ultimate creep coefficient of (∆t)Ft + G is a camber; and
analysis using Ect and Ig.
• Ft = initial prestressing force
concrete;
η = F/Ft; immediately after transfer.
kr = 1/(1 + As/Aps); and
KCA = age at loading factor for creep
• Uncracked section;
• The pressure line is assumed
The long-term deflection is estimated resulting from the sustained
from: loadings;
2 2
l l • The profile of the pressure
∆ ( t ) = φ 1 ( t ) --- + [ φ 2 ( t ) – φ 1 ( t ) ] ------
∆t is obtained using Ig and the 8 48 line is assumed parabolic;
Naaman (1982 and predicted elastic modulus at where • Prestress losses must be
1983) time of loading Ec (t). φ1(t) = midspan curvature at time t; estimated a priori;
φ2(t) = support curvature at time t; • Design chart is provided for
φ(t) = M/[Ece(t) × I]; and the equivalent modulus; and
Ece(t) = equivalent modulus • As and A′s are accounted for
through It and neutral axis of
bending.
For cracked members, the
short-term deflection is Long-term deflection is not addressed
Bronson and Trost but it is assumed that for a given ∆t, the
(1982) computed using Ieff modified • Cracked members.
for partial prestressing. earlier method is applicable.

∆ add = λ 1 ( ∆ i ) G + λ 2 ( ∆ i ) F • kr = same as Branson;


i

E ci • Uncracked section;
∆t is obtained from elastic λ 1 = k r ------α • Design values of λ1 and λ2
Martin (1977) analysis using Ect and Ig. Ec
were recommended; and
α = ( 2 – 1.2A s ′ ⁄ A s ) ≥ 0.6 • The method is adopted in
λ 2 = ηλ 1 PCI Design Handbook.

The long-term deflection is obtained • Uncracked sections; and


∆t is obtained from elastic by integrating the curvatures modified • For common loading cases,
Tadros et al. (1975 and
1977) by a creep recovery parameter and a only the curveatures at the
analysis using Ec(t) and Ig. relaxation reduction factor that are support and midspan
time-dependent. sections are needed.
The long-term deflection is obtained
by integrating the curvature along the
member. The time-dependent
curvature is modified by the effect of
∆t is obtained from long-term an equivalent force acting at the
deflection expression at initial centroid of the prestressing steel due
to creep and shrinkage strain. • Uncracked sections; and
Dilger (1982) loading time. The age adjusted • A relaxation reduction factor
effective modulus and a creep Mc
φ ( t ) = φ i C c ( t ) – -------------------
- is used.
transformed moment of inertia I tr E ca ( t )
are used.
Itr = transformed moment of inertia;
Mc = moment due to equivalent
transformed force; and
Eca(t) = age adjusted modulus
The nonlinear analysis takes both • A computer program is
∆t is obtained from nonlinear
creep and shrinkage into account, available from PCI to
Moustafa (1986) analysis using actual material using ACI creep and shrinkage perform the nonlinear
properties. functions and a time step method. analysis.
423.5R-10 ACI COMMITTEE REPORT

Fig. 2.4—Equivalent loads and moments produced by prestressing tendons (Nilson 1987);
P = prestressing force.

prestressing, as well. This approach allows the designer to cracking, or as the capacity of an equivalent member without
treat a prestressed concrete member as if it was a transverse reinforcement. Therefore, detailed expressions
nonprestressed concrete member. The prestressing steel is have been developed in terms of parameters relevant to the
treated as mild (passive) reinforcement for ultimate strength of members without transverse reinforcement.
conditions, with a remaining tensile capacity of (fps – fpe), These parameters include the influence of axial compres-
where fps is the stress in the reinforcement at nominal sion, member geometry, support conditions, axial tension,
strength, and fpe is the effective stress in the prestressed and prestress.
reinforcement (after allowance for all losses). 2.5.2 Shear—The following behavioral changes occur in
Most codes of practice (ACI 318; AASHTO Bridge partially prestressed members at nominal shear levels, as
Design Specifications, Eurocode 2; and CSA Design of Con- some of the longitudinal prestressing steel in the tension face
crete Structures for Buildings) use sectional methods for of the member is replaced by mild reinforcement, but the
design of conventional beams under bending, shear, and tor- same total flexural strength is maintained:
sion. Truss models provide the basis for these sectional • Due to the lower effective prestress, the external load
design procedures that often include a term for the concrete required to produce inclined cracking is reduced. This
contribution (Ramirez and Breen 1991). The concrete contri- results in an earlier mobilization of the shear reinforce-
ment; and
bution supplements the sectional truss model to reflect test
• After inclined cracking, there is a reduction in the con-
results in beams and slabs with little or no shear reinforce-
crete contribution. The reduction is less significant as the
ment and to ensure economy in the practical design of such degree of prestressing decreases. This can be explained
members. as follows:
In design specifications, the concrete contribution has •The addition of mild reinforcement results in an
been taken as either the shear force or torsional moment at increase in the cross-sectional area of the longitudinal
PARTIALLY PRESTRESSED CONCRETE 423.5R-11

tension reinforcement and the reinforcement stiffness; supported by the theory of plasticity, in which a space truss
and with variable inclination of compression diagonals provides
•The increase in stiffness of the longitudinal tension a lower-bound (static) solution.
reinforcement delays the development of the crack- This procedure is representative of the behavior of thin-
ing pattern, so that the cracks are narrower and the walled tubes in torsion. For these members, the shear
flexural compression zone is larger than in fully pre- stresses induced by torsion can be determined using only
stressed members of comparable flexural strength.
equilibrium relationships. Because the wall of the tube is
These behavioral changes are well documented in a series thin, a constant shear stress can be assumed across its thick-
of shear tests carried out by Caflisch et al. (1971). In this ness. In the longitudinal direction, equilibrium conditions
series of tests, the only variable was the degree of prestress- dictate that the torsion-induced shear stresses be resisted by
ing. The cross sections of the prestressing steel and the rein- a constant shear flow around the perimeter of the section.
forcing steel were selected so that all the beams had the same For other sections before cracking, the strength in torsion
flexural strength. These tests also showed that for the same can be computed from the elastic theory (de Saint-Venant
external load, a higher degree of prestressing delays the 1956) or from the plastic theory (Nadai 1950). Rather than
onset of diagonal cracking and results in a decrease in the using these more complex approaches, an approximate pro-
stirrup forces. The decrease in stirrup forces can be cedure is used in ACI 318 based on the concept that most tor-
explained by the fact that a higher degree of prestressing in sion is resisted by the high shear stresses near the outer
the web of the member results in a lower angle of inclination perimeter of the section. In this approach, the actual cross
of the diagonal cracks. The lower angle of inclination of the section before cracking is represented by an equivalent thin-
cracks leads to the mobilization of a larger number of stir- walled tube with a wall thickness t of
rups.
In ACI 318, a cursory review of the design approach for
A cp
shear indicates that partially prestressed members can be t = 0.75 -------
- (2-2)
designed following the same procedure as for fully pre- P cp
stressed members. In ACI 318, it is assumed that flexure and
shear can be handled separately for the worst combination of where Acp = area enclosed by outside perimeter of concrete
flexure and shear at a given section. The analysis of a beam cross section, and Pcp = outside perimeter of the concrete
under bending and shear using the truss approach clearly cross section. While the area enclosed by the shear flow
indicates that, to resist shear, the member needs both stirrups path, Ao, could be calculated from the external dimensions
and longitudinal reinforcement. The additional longitudinal and wall thickness of the equivalent tube, it is reasonable to
tension force due to shear can be determined from equilibri- approximate it as equal to 2Acp/3. Cracking is assumed to
um conditions of the truss model as (V cot θ), where V is the occur when the principal tensile stress reaches 4√f c′ . For pre-
shear force at the section, and θ is the angle of inclination of stressed members, the cracking torque is increased by the
the inclined struts with respect to the longitudinal axis of the prestress. A Mohr’s Circle analysis based on average stress-
member. es indicates that the torque required to cause a principal ten-
In the shear provisions of ACI 318, no explicit check of the sile stress equal to 4√f c′ is the corresponding cracking torque
shear-induced force in the longitudinal reinforcement is per- of a nonprestressed beam times
formed (Ramirez 1994). The difference between the flexural
strength requirements for the prestress reinforcement and the f pc
ultimate tensile capacity of the reinforcement can be used to 1 + ------------
- (2-3)
satisfy the longitudinal tension requirement. The 1994 AASH- 4 f c′
TO LRFD Bridge Design Specifications, in the section for
shear design, includes a check for longitudinal reinforcement. where fpc in psi is the average precompression due to pre-
These recommendations are based on a modified compres- stress at the centroid of the cross section resisting the exter-
sion field theory (Vecchio and Collins 1986). nally applied loads or at the junction of web and flange if the
2.5.3 Torsion—ACI 318 includes design recommenda- centroid lies within the flange.
tion for the case of torsion or combined shear and torsion in In ACI 318, the design approach for combined actions
prestressed concrete members. These provisions model the does not explicitly consider the change in conditions from
behavior of a prestressed concrete member before cracking one side of the beam to the other. Instead, it considers the
as a thin-walled tube and after cracking using a space-truss side of the beam where shear and torsional effects are addi-
model with compression diagonals inclined at an angle θ tive. After diagonal cracking, the concrete contribution of
around all faces of the member. For prestressed members, θ the shear strength Vc remains constant at the value it has
can be taken equal to 37.5 degrees if the effective prestress- when there is no torsion, and the torsion carried by the con-
ing force is not less than 40% of the tensile strength of the crete is taken as zero. The approach in ACI 318 has been
prestressed reinforcement. For other cases, θ can be taken compared with test results by MacGregor and Ghoneim
equal to 45 degrees. This approach is based on the work car- (1995).
ried out in the 1960s and 1970s by European investigators In the AASHTO LRFD Specifications, the modified com-
led by Thurlimann (1979). This work proposed a method pression field theory proposed for members under shear has
423.5R-12 ACI COMMITTEE REPORT

components, which are generally grouped into instantaneous


and time-dependent losses.
Instantaneous losses are due to elastic shortening, anchorage
seating, and friction. They can be computed for partially pre-
stressed concrete in a manner similar to that for fully pre-
stressed concrete.
Time-dependent losses are due to shrinkage and creep of
concrete and relaxation of prestressing steel. Several methods
are available to determine time-dependent losses in pre-
stressed concrete members: lump-sum estimate of total loss-
es, lump-sum estimates of separate losses (such as loss due
to shrinkage or creep), and calculation of losses by the time-
step method. Because the numerical expressions—
described, for instance, in the AASHTO’s Standard Specifi-
cations for Highway Bridges or the PCI Design Handbook
Fig. 3.1—Typical stress change in prestressing steel at end (1992)—for the first two methods were developed assuming
of service life for sections cracked and uncracked under full prestressing, they are not applicable to partially pre-
permanent loads (Watcharaumnuay and Naaman 1985). stressed members.
been extended to include the effects of torsion. Similar to the The combined presence of nonprestressed reinforcing bars
ACI 318 procedure, the AASHTO approach concentrates on and a lower level of prestress in partially prestressed concrete
the design of the side of the beam where the shear and tor- should lead to smaller time-dependent prestress losses than
sional stresses are additive. for fully prestressed concrete. Another significant factor is
that a partially prestressed member can be designed as a
As in the case of shear, torsion leads to an increase in the
cracked member under sustained loading. A computerized
tensile force on the longitudinal reinforcement. The longitu-
time-step analysis of the concrete and steel stresses along par-
dinal reinforcement requirement for torsion should be super-
tially prestressed sections shows that the effect of creep of
imposed with the longitudinal reinforcement requirement for
concrete on the stress redistribution between the concrete and
bending that acts simultaneously with the torsion. In ACI steel tends to counteract the effect of prestress losses over
318, the longitudinal tension due to torsion can be reduced time. This is particularly significant for members that are
by the compressive force in the flexural compression zone of cracked under permanent loads. The result is illustrated in
the member. Furthermore, in prestressed beams, the total Fig. 3.1 (Watcharaumnuay and Naaman 1985), which was
longitudinal reinforcement, including tendons at each sec- derived from the analysis of 132 beams with various values
tion, can be used to resist the factored bending moment plus of the partially prestressed ratio (PPR) and the reinforce-
the additional tension induced by torsion at that section. ment index ω. While time-dependent prestress losses can be
ACI 318 and AASHTO Specifications recognize that, in 14% for uncracked fully prestressed sections, they remain
many statically indeterminate structures, the magnitude of low for cracked sections up to relatively high values of the
the torsional moment in a given member will depend on its PPR. Fig. 3.1 also shows that for uncracked sections, time-
torsional stiffness. Tests have shown (Hsu 1968) that when a dependent prestress losses decrease with a decrease in PPR.
member cracks in torsion, its torsional stiffness immediately Creep redistribution of force to reinforcing steel may reduce
after cracking drops to approximately 1/5 of the value before the precompression in the concrete.
cracking, and at failure can be as low as 1/16 of the value An investigation under NCHRP Project 12-33 has
before cracking. This drastic drop in torsional stiffness addressed prestress losses in partially prestressed normal and
allows a significant redistribution of torsion in certain high-strength concrete beams. This work was described by
indeterminate beam systems. In recognizing the reduction of Naaman and Hamza (1993) and was adopted in the Final
torsional moment that will take place after cracking in the Draft LRFD Specifications for Highway Bridge Design and
case of indeterminate members subjected to compatibility- Commentary (Transportation Research Board 1993). It led to
induced torsion, ACI 318R states that a maximum factored lump-sum estimates of time-dependent losses for partially
torsional moment equal to the cracking torque can be prestressed beams that are assumed uncracked under the
assumed to occur at the critical sections near the faces of design sustained loading. These are summarized in Table 3.1
supports. This limit has been established to control the width and can be used as a first approximation in design.
of the torsional cracks at service loads.
3.2—Cracking
CHAPTER 3—TIME-DEPENDENT BEHAVIOR Crack widths in reinforced and cracked partially
3.1—Prestress losses prestressed concrete members subjected to sustained loads
The stress in the tendons of prestressed concrete structures are known to increase with time. Bennett and Lee (1985)
decreases continuously with time. The total reduction in reported that crack widths increase at a fast rate during the
stress during the life span of the structure is termed “total early stages of loading then tend toward a slow steady rate of
prestress loss.” The total prestress loss consists of several increase. This is not surprising because deflections and
PARTIALLY PRESTRESSED CONCRETE 423.5R-13

Table 3.1—Time-dependent losses, ksi (LRFD Bridge Design Specifications


1994)
Type of beam For wires and strands with
section Level fpu = 235, 250, or 270 ksi For bars with fpu = 145 or 160 ksi
Rectangular beams Upper bound 29.0 + 4.0 PPR
19.0 + 6.0 PPR
and solid slab Average 26.0 + 4.0 PPR
Box girder Upper bound 21.0 + 4.0 PPR
15.0
Average 19.0 + 4.0 PPR
f c′ – 6.0
I-girder Average 33.0 1.0 – 0.15 ------------------
- + 6.0PPR 19.0 + 6.0 PPR
6.0

f c′ – 6.0
Upper bound 39.0 1.0 – 0.15 ------------------
- + 6.0PPR
Single-T, double-T, 6.0 f c′ – 6.0
hollow core, and 31.0 1.0 – 0.15 ------------------
- + 6.0PPR
voided slab f c′ – 6.0 6.0
Average 33.0 1.0 – 0.15 ------------------
- + 6.0PPR
6.0

camber increase with time. Crack widths, however, represent


only localized effects, and their relative increase with time is
not proportional to the increase of deflections.
No studies have been reported where an analytical model
of crack width increase with time was developed. An inves-
tigation by Harajli and Naaman (1989), however, discussed
in Chapter 4 of this report, has led to the development of a
model to predict the increase in crack width under cyclic
fatigue loading. The model accounts for the effect of change
in steel stress due to cyclic creep of concrete in compression,
the increase in slip due to bond redistribution, and concrete
shrinkage.

3.3—Deflections
Several experimental investigations have dealt with the
time-dependent deflection of partially prestressed concrete
members (Bennett and Lee 1985; Bruggeling 1977; Jittawait
and Tadros 1979; Lambotte and Van Nieuwenburg 1986;
Abeles 1965; and Watcharaumnuay and Naaman 1985).
Deflections and cambers in partially prestressed concrete
members are expected to vary with time similarly to rein-
forced or fully prestressed concrete. When positive deflec- Fig. 3.2—Long-term deflections of fully prestressed and
tion (opposite to camber) is present, in all cases the partially prestressed (cracked and uncracked) beams
deflection in partially prestressed beams falls between those (Naaman 1982); 1 in. = 25.4 mm.
of reinforced concrete and fully prestressed concrete beams was about 1.25 for pretensioned members and 1.5 for post-
(Fig. 3.2). A limited experimental study by Jittawait and tensioned members.
Tadros (1979) also seems to confirm this observation. Several analytical investigations have dealt with the time-
A study of the long-term behavior of partially prestressed dependent deflections of prestressed and partially
beams has been conducted at the Magnel laboratory in Bel- prestressed beams assumed to be uncracked (Table 2.2). The
gium (Lambotte and Van Nieuwenburg 1986). Twelve par- evaluation of deflections for cracked, partially prestressed
tially prestressed beams with PPR of 0.8, 0.65, and 0.5 were members has been conducted by several researchers
either kept unloaded or were loaded with an equivalent full (Branson and Shaikh 1985; Ghali and Tadros 1985; Tadros
service load. For the unloaded beams, increased camber with et al. 1985; Ghali and Favre 1986; Al-Zaid et al. 1988;
time was generally observed for PPR = 0.8; for PPR = 0.65, Elbadry and Ghali 1989; Ghali 1989; and Founas 1989).
the camber reached a peak value and then decreased with Watcharaumnuay and Naaman (1985) proposed a method to
time, resulting in a practically level beam; and for PPR = 0.5, determine time- and cyclic-dependent deflections in simply
the initial camber decreased with time, resulting in a final supported, partially prestressed beams in both the cracked
downward deflection. For the loaded beams, deflections and uncracked state. The time-dependent deflection is
were observed in all cases and increased with time. These treated as a special case of cyclic deflection. Compared with
observations are illustrated in Fig. 3.3. After 2 years of load- the time-step method where the deflection is obtained from
ing, the ratio of additional deflection to the initial deflection the summation of deflection increments over several time
423.5R-14 ACI COMMITTEE REPORT

KD , KF = constants depending on type of loading and


steel profile;
M = sustained external moment at midspan;
F(t) = prestressing force at time t;
Ieff (t) = effective moment of inertia of cracked section
at time t; and
Ece (t) = equivalent elastic modulus of concrete at time t.
The eqivalent elastic modulus of concrete can be approxi-
mated by the following equation

Ec ( t )
Fig. 3.3(a)—Variation with time of midspan deflection for E ce ( t ) = -------------------------------- (3-2)
unloaded specimens (Lambotte and Van Nieuwenburg 1 + Cc ( t – tA )
1986); 1 in. = 25.4 mm.
where
t = time or age of concrete;
tA = age of concrete at time of loading;
Ec (t) = instantaneous elastic modulus of concrete at
time t;
and
Cc (t - tA) = creep coefficient of concrete at time t when
loaded at time tA.
Several expressions are available to predict the creep coef-
ficient of concrete. The recommendations of ACI Committee
209 (1982) can be followed in most common applications.
Fig. 3.3(b)—Variation with time of midspan deflection for
specimens under service loading (Lambotte and Van 3.4—Corrosion
Nieuwenburg 1986); 1 in. = 25.4 mm. The reinforcement in a fully prestressed member is better
protected against corrosion than the reinforcement in a
intervals, this method leads to the deflection at any time t and partially prestressed member. Cracks in partially prestressed
cycle N directly. The method satisfies equilibrium and strain beams are potential paths for the passage of corrosive agents.
compatibility. Using a slightly different approach, the Although corrosion also occurs along uncracked sections,
method proposed by Watcharaumnuay and Naaman (1985) cracking can facilitate corrosion. Abeles (1945) suggested
was generalized by Al-Zaid et al. (1988) and Founas (1989), that corrosion of the prestressing steel in partially prestressed
and was extended to include composite beams as well as members can be mitigated by requiring that the member
noncomposite beams. faces that are cracked under full service load be in
In dealing with time-dependent deflections, several time- compression under permanent (dead) loads. He demonstrated
the effectiveness of this strategy in the behavior of beams
dependent variables should be determined. These include the
partially prestressed using small-diameter wires that were used
prestressing force, the moment of inertia of the section, and
in the roof of an engine shed for steam locomotives. These
the equivalent modulus of elasticity of the concrete. The
beams successfully resisted a very corrosive atmosphere caused
expression for the effective moment of inertia described in by the mixture of smoke and steam ejected onto them from the
Eq. (2-1) can be used. In this expression, however, Mcr and funnels of the locomotives.
Icr are time-dependent variables because they depend on the Limiting the size of crack widths to reduce the probability
value of the prestressing force and the location of the neutral of corrosion has been common practice in design. Later stud-
axis (zero stress point along the section), both of which vary ies (ACI 222R-89), however, move away from this approach
with time. The equivalent modulus of elasticity of the con- by pointing out that corrosion is due to many causes, most of
crete depends on the variation of creep strain with time. which can proceed with or without cracking to be activated.
The following method can be used to estimate deflection Corrosion in prestressing steels is much more serious than
at any time t in a cracked, simply supported beam corrosion in nonprestressed reinforcing steels. Prestressing
steel is generally stressed to over 50% of its strength, making
KD M KF F ( t ) it susceptible to stress corrosion, and the diameter of individ-
∆ ( t ) = ----------------------------
- + ----------------------------
- (3-1) ual prestressing steel wires is relatively small. Even a small,
E ce ( t )I eff ( t ) E ce ( t )I eff ( t )
uniform corrosive layer or a corroded spot can progressively
reduce the cross-sectional area of the steel and lead to wire
where failure.
PARTIALLY PRESTRESSED CONCRETE 423.5R-15

Corrosion is mostly an electrochemical problem and increased crack widths and deflections under service loads
should be treated accordingly. Precautions should be taken (Naaman 1982b; Shakawi and Batchelor 1986; and ACI
to prevent or to reduce prestressing steel corrosion. Committee 215 1974). The increase in crack widths in par-
ACI 423.3R addresses the historical causes of corrosion in tially prestressed beams, however, has been smaller than
unbonded tendons. The Post-Tensioning Manual (Post-Ten- that generated in similarly loaded, precracked, fully pre-
sioning Institute 1990) provides guidance for corrosion pro- stressed beams (Harajli and Naaman 1984).
tection for bonded and unbonded tendons. Occasionally, for Because the proportion of dead to total load often increas-
pretensioned concrete, epoxy-coated prestressing strands es as the span length increases, the significance of fatigue as
have been specified for corrosive environments. High curing a critical limit state tends to diminish as span lengths
temperatures, however, could adversely affect the bond of increase (Freyermuth 1985).
epoxy-coated strand. Guidelines for the Use of Epoxy-Coat- Abeles demonstrated the practicability of using partially
ed Strand (PCI Ad Hoc Committee 1993) contains recom- prestressed concrete members when fatigue resistance is a
mendations for its use. serious consideration (Abeles 1954). He persuaded British
Lenschow (1986) reported that crack widths less than Railways to consider the use of partial prestressing in the
0.004 to 0.006 in. (0.1 to 0.15 mm), which develop under reconstruction of highway bridges over the London to
maximum load, will heal under long-term compression. Manchester line, when it was electrified around 1950. Brit-
Crack widths that increase to less than 0.01 in. (0.3 mm) ish Railways financed extensive cyclic loading tests of full-
under rare overload (every 1 to 3 years) can reduce to 0.004 scale members, which were designed to allow 550 psi
to 0.006 in. (0.1 to 0.15 mm) under sustained compression. (approximately 8√f c′ or 3.8 MPa) tension under full service
Keeping crack widths under such limits should avoid prob- load and 50 psi (0.34 MPa) compression under dead load
lems with corrosion. only. These members were cracked under static load and
were then subjected to 3 million cycles of load producing the
CHAPTER 4—EFFECTS OF REPEATED design range of stress, 50 psi (0.34 MPa) compression to 550
LOADING (FATIGUE) psi (3.8 MPa) tension at the flexural tension face. Behavior
4.1—Background was satisfactory, with essentially complete closure of cracks
Two major requirements should be considered when and recovery of deflection after 3 million cycles of load. The
designing members subjected to repeated loads: member strength under static loading was not decreased by the cyclic
strength and serviceability. The static strength and fatigue loading. Many relatively short-span bridges were construct-
strength of the member should exceed loads imposed and ed using such partially prestressed members and they per-
adequate serviceability requirements (deflection and crack formed satisfactorily.
control) should be provided. Recently, Roller et al. (1995) conducted an experimental
The fatigue strength of a member is affected primarily by program including four full-size, pretensioned, bulb-tee gird-
the stress range (difference between maximum and mini- ers made with high-strength concrete and pretensioned. The
mum stress), the number of load applications or cycles, and girders were 70 ft (21.3 m) long and 54 in. (1.4 m) deep with
the applied stress levels. The fatigue life of a member is a concrete compressive strength of 10,000 psi (69 MPa). One
defined as the number of load cycles before failure. The of the four test girders with a simple span of 69 ft (21.0 m)
higher the stress range imposed on the member, the shorter was subjected to cyclic (fatigue) flexural loading using two
the fatigue life. point loads spaced 12 ft (3.66 m) apart at midspan. A con-
Reliability analyses indicate that the probability of fatigue crete deck 10 ft (3.05 m) wide and 9.5 in. (250 mm) thick had
failure of reinforcement in partially prestressed beams is been cast on the girder to represent the effective flange of the
higher on average than failure by any other common service- composite girder in a bridge.
ability or ultimate limit state criterion (Naaman 1985, Naa- During the cyclic flexural loading, the upper limit of the
man and Siriaksorn 1982). Fatigue can be a critical loading load produced a midspan tensile stress at the extreme fiber of
condition for partially prestressed concrete beams because the lower flange equal to 6√f c′ . The lower limit of the load was
high stress ranges can be imposed on the member in the ser- selected such that a steel stress range of 10,000 psi (69 MPa)
vice-load range (Naaman and Siriaksorn 1979). would be produced. After each million cycles of loading, the
Partially prestressed concrete beams generally crack upon girder was tested statically to determine its stiffness. Slight
first application of live load (Naaman 1982b). Subsequent reductions in stiffness and camber were observed, but there
applications of live load cause the cracks to reopen at the was no significant change in prestress loss. The girder per-
decompression load (when the stress at the extreme tensile formed satisfactorily for 5 million cycles of fatigue loading.
face is zero), which is less than the load that caused first After completion of the long-term fatigue load test, the gird-
cracking. To maintain equilibrium in the section after crack- er was tested under static load to determine its ultimate flex-
ing, the neutral axis shifts toward the extreme compression ural strength. It developed an ultimate moment equal to 94%
fiber. This shift generates higher strains (stresses) in the tensile of the ultimate moment capacity of a companion girder that
reinforcement. had been under long-term sustained load. The measured
Under repeated loads, the larger stress changes (created moment capacity also exceeded the calculated moment
by opening and closing of the cracks) cause fatigue damage capacity by 7.5% based on the AASHTO Standard Specifi-
in the constituent materials, bond deterioration, and cations for Highway Bridges.
423.5R-16 ACI COMMITTEE REPORT

Tests have been conducted on ordinary reinforcement, both


in-air and embedded in concrete, to determine its fatigue prop-
erties. These tests have yielded varying results (Rehm 1960,
Soretz 1965). For straight deformed bars, ACI Committee 215
(1974), Model Code for Concrete Structures (CEB-FIP 1978),
FIP Commission on Model Code (1984), and Ontario High-
way Bridge Design Code (Ministry of Transportation and
Communications 1983) recommend stress range limits of 20,
22, and 18 ksi (138, 152, and 124 MPa), respectively. The
lowest stress range found to cause fatigue failure in a hot-
rolled bar is 21 ksi (145 MPa) (ACI Committee 215 1974).
ACI 343R recommends limiting the reinforcement stress
range in terms of the minimum stress and reinforcement
deformation geometry

f f = 21 – 0.33f min + 8 ( r ⁄ h ) (4-2)

where
ff = safe stress range, ksi;
fmin = minimum applied stress, ksi; and
r/h = ratio of base radius-to-height of rolled-on trans-
verse deformation (a value of 0.3 can be used in
the absence of specific data).
The fatigue strength of prestressing reinforcement
Fig. 4.1—Comparison of observed fatigue life of prestressing
strands with existing data (Harajli and Naaman 1985a). depends upon the steel type (bar, wire, strand), anchorage
(unbonded post-tensioned reinforcement), extent of bond (ACI
Committee 215 1974), and steel treatment. Paulson et al.
4.2—Material fatigue strength (1983) conducted fatigue tests (in-air) of 50 seven-wire
The fatigue resistance of a structural concrete member is strand samples obtained from six different manufacturers.
directly related to the fatigue properties of its component All of the strands conformed with ASTM A 416 require-
materials (Naaman 1982b). Therefore, the fatigue behavior ments. The minimum stresses applied in the tests ranged
of the constituent materials should be investigated first. from 75 to 165 ksi (517 to 1138 MPa), and the stress ranges
Fatigue of concrete—The applied stress range limit for varied from 22 to 81 ksi (152 to 559 MPa). A significant
concrete recommended by ACI 215 (1974) is given by the variation was observed in results from even two samples of
formula the same product produced by the same manufacturer. The
effect of the end grips dominated the fatigue curves in the
f cr = 0.4f c′ – f min ⁄ 2 (4-1) region of long-life, low-stress-range.
The following relationship was found to lie above 95 to
where 97.5% of the failure points
fcr = maximum recommended stress range for concrete;
f c′ = specified concrete compressive strength; and log N = 11 – 3.5 log f sr (4-3)
fmin = minimum applied stress.
Concrete can sustain a fluctuating stress between zero where
and 50% of its static strength for approximately 10 million N = number of cycles; and
cycles in direct compression, tension, or flexure without fsr = maximum stress range for a fatigue life of N cy-
failure (Norby 1958; Gylltoft 1978; McCall 1958; Stelson cles, ksi.
and Cernica 1958; and Hilsdorf and Kesler 1966). Concrete The researchers did not find the effect of minimum stress on
stresses resulting from service loads are generally smaller fatigue life great enough to warrant inclusion in the equation.
than this magnitude (Shahawi and Batchelor 1986). Conse- The FIP Commission on Prestressing Steel (1976) recom-
quently, concrete fatigue failure generally will not control in mends a stress range of 15% of fpu with a minimum applied
the case of repetitively loaded partially prestressed beams stress not greater than 75% of fpu for a fatigue life of 2 mil-
(Naaman 1982b; Harajli and Naaman 1984; and Bennett lion cycles. For the same fatigue life of two million cycles,
1986). however, Naaman (1982b) recommends a reduced stress
Fatigue of reinforcement—Fatigue failure of underreinforced range of 10% fpu with a minimum applied stress not greater
prestressed concrete beams is believed to be governed by the than 60% of fpu to better correlate with test results (Fig. 4.1).
fatigue failure of the steel reinforcement (Warner and Huls- The following equation can be used to predict other maxi-
bos 1966a). mum safe stress ranges
PARTIALLY PRESTRESSED CONCRETE 423.5R-17

f sr ⁄ f pu = – 0.123 log N f + 0.87 (4-4)

where
fsr = maximum safe stress range for a fatigue life of N
cycles;
fpu = specified tensile strength of the prestressing
strand; and
Nf = number of cycles to failure.
The endurance limit (stress range for which the reinforce-
ment will not fail for an infinite number of cycles) has not
been found for prestressing steel (Naaman 1982b); however,
a fatigue life of 2 million cycles is considered to be sufficient
for most applications.
The previous discussion applies to pretensioned strands. Fig. 4.2—Fatigue resistance of post-tensioned tendon in
steel duct and in thick-walled plastic duct (Oertle 1988);
For post-tensioned tendons, two more levels of fatigue 1 ksi = 6.9 MPa.
strength have to be considered: the strand/duct assembly and
the tendon anchorages. For the strand/duct assembly, fretting live load (Fig. 4.3). For the same type of beam section, the
fatigue may govern if high contact stresses between strand effect of the PPR was plotted with respect to the reinforce-
and corrugated steel duct are combined with small relative ment stress range due to the application of live loads (Fig.
movements at cracks. Under such circumstances, the fatigue 4.4). The discontinuity in the plots corresponds with first
strength of the strand/duct assembly can drop to as low as cracking of the concrete in the beams. It is evident from the
14,300 psi (100 MPa). Fatigue strengths of anchorages are in figures that higher stress ranges are associated with partially
the order of 14,300 psi (100 MPa), according to FIP Com- prestressed sections. Thus, fatigue problems are more signif-
mission on Prestressing Steel and Systems (1992). icant in partially prestressed sections than in their ordinary
Designers typically place tendon anchorages away from reinforced or fully prestressed counterparts.
areas with high stress variations and avoid fatigue problems
at the anchorages. A similar approach normally will not 4.4—Prediction of fatigue strength
work to avoid fretting fatigue because maximum stresses The studies described have a common conclusion summa-
often occur at sections with maximum tendon curvature and rized by Naaman (1982b) and Warner and Hulsbos (1966b).
maximum contact stresses between strand and duct. Fretting The critical limit state (fatigue failure) of partially pre-
fatigue between strand and duct, however, can be avoided by stressed concrete beams is generally due to failure of the
using thick-walled plastic ducts rather than corrugated steel reinforcement. The fatigue life of the member can be predict-
ducts (Oertle 1988). With a thick-walled plastic duct, the ed from the smaller of the fatigue lives of the reinforcing
strand reaches fatigue strengths comparable to those of steel or the prestressing steel. Many of these investigations
strand in air. Fig. 4.2 shows the fatigue performance of ten- have indicated that in-air test results of reinforcement pro-
dons with steel and plastic ducts in simply supported beams vide a good indication of the member fatigue life.
under four-point loading. In the specimen with a steel duct, Naaman therefore recommends using Eq. (4-4) or Fig. 4.1
50% of the tendons failed at a fatigue amplitude of 25,000 psi to estimate the fatigue life of stress-relieved seven-wire
(175 MPa); in contrast, only 18% of the tendons in the spec- strand for the appropriate stress range. A strand subjected to
imen with a plastic duct failed at a fatigue amplitude of a minimum stress less than 60% of its tensile strength with a
39,400 psi (275 MPa). stress range of 10% of the tensile strength should provide a
fatigue life of approximately two million cycles.
4.3—Fatigue in partially prestressed beams For ordinary reinforcement, Naaman recommends using
To illustrate the relative importance of fatigue for partially Eq. (4-2) to determine safe stress ranges that provide fatigue
prestressed beams compared with that for ordinary reinforced lives in excess of 2 million cycles.
or fully prestressed beams, Naaman (1982b) analyzed three ACI Committee 215 (1974) and Venuti (1965) recom-
concrete beams, identical except for the partially prestressed mend conducting a statistical investigation of at least six to
reinforcement ratio (PPR = 0, 0.72 and 1.0). Note that PPR 12 reinforcement samples at appropriate stress levels to estab-
= 0 represents an ordinary reinforced beam; PPR = 1.0 rep- lish the fatigue characteristics of the material. At least three
resents a fully prestressed beam; and PPR = 0.72 represents stress levels are required to establish the finite-life portion of
a partially prestressed beam. All of the beams were designed the S-N diagram: one stress level near the static strength, one
to provide the same ultimate moment capacity. Material near the fatigue limit, and one in between.
properties and relevant data are given by Naaman and Siri- The choice of the PPR and relative placement of the reinforce-
aksorn (1979). ment have a significant effect on the fatigue response of the
For each beam, computed stress ranges in ordinary and pre- members. Naaman (1982b) states that proper selection of these
stressed steel were plotted with respect to the applied load (in variables can maintain the stress ranges in the reinforcement to
excess of the dead load) varying from zero to the specified within acceptable limits.
423.5R-18 ACI COMMITTEE REPORT

Fig. 4.3—Typical comparison of stress changes in steel for reinforced, prestressed, and
partially prestressed beams (Naaman 1982a).

Balaguru (1981) and Balaguru and Shah (1982) have present-


ed a method and a numerical example for predicting the fatigue
serviceability of partially prestressed members. The method
compares the stress ranges in the beam constituents to the fatigue
limits of each individual component (concrete, prestressing steel
and nonprestressing steel) using the equations derived by Naa-
man and Siriaksorn (1979) to calculate stresses for both
uncracked and cracked sections.
Naaman and Founas (1991) also presented models to cal-
culate the structural responses that account for shrinkage,
static and cyclic creep, and relaxation of prestressing steel.
For any time t and cycle N, the models can be used to com-
pute stresses, strains, curvatures, and deflections.

4.5—Serviceability aspects
In a cracked concrete member, whether nonprestressed,
partially prestressed, or fully prestressed, the crack widths
and deflections generally increase under repeated loadings
(Naaman 1982b).
The increase in crack widths and deflections in concrete
members is mostly attributed to the cyclic creep of concrete
and bond deterioration accompanied by slip between the
reinforcement and concrete on either side of existing cracks.
Fig. 4.4—Typical stress changes in steel at different levels of ACI Committee 224 (1980) notes that 1 million cycles of
prestressing (Naaman 1982a). load can double the crack widths.
PARTIALLY PRESTRESSED CONCRETE 423.5R-19

Harajli and Naaman (1989) developed an analytical model


for cyclic slip and stresses to compute crack widths in par-
tially prestressed concrete beams. In the analysis, equilibri-
um was assumed between the reinforcement bond stresses
and the concrete tensile stresses in a concrete tensile prism
joining two successive cracks. In addition, a local bond
stress-slip relationship was assumed for the reinforcement.
The results of the analyses were in agreement compared
with the experimental results shown in Fig. 4.5 and 4.6 for
monotonic and cyclic loading, respectively. Agreement was
also obtained with tests conducted by Lovegrove and El Din
(1982) for which the steel stresses ranged from 7 to 43 ksi
(48 to 297 MPa).
The analytical models were also useful in predicting trends
in the crack patterns in terms of growth and mechanisms of
growth. From the analyses, Harajli and Naaman (1989)
attributed increased crack widths to increases in steel
stresses caused by cyclic creep of concrete in compression
and bond redistribution between cracks.
They indicated that partially prestressed beams have
smaller crack spacings and less crack growth than fully pre-
stressed beams subjected to fatigue loads. The presence of Fig. 4.5—Comparison of observed and predicted crack
ordinary reinforcement in prestressed beams helps to control widths for monotonically tested beams (Harajli and
cracking in static and fatigue tests. Ordinary reinforcement Naaman 1989); 1 ksi = 6.9 MPa; 1 in. = 25.4 mm.
reduces the increased steel stresses attributed to cyclic creep
of concrete. In addition, it minimizes bond redistribution. Tests by Shahawi and Batchelor (1986) indicated a similar
Crack widths were observed to be a function of reinforce- linear relationship between crack width and reinforcement
ment stress and crack spacing. In tests that generated similar stress. They recommend the use of the FIP-CEB equation
crack spacings (3 to 4.5 in., [76 to 114 mm]), a nearly linear (Eq. 4-8) for predicting the maximum crack widths in par-
relationship was observed between crack widths and reinforce- tially prestressed beams subjected to fatigue
ment stress (Fig. 4.7). To estimate the crack spacing acs and
crack width W the following equations were given W max = f s × 10
–3
(4-8)

a cs = 1.20A t ⁄ P (4-5)
where
Wmax = maximum crack width, mm; and
W = 2S o B (4-6) fs = change in steel stress from decompression at the
extreme tensile fiber, MPa.
W N = 2S o, N B N + [ ( f so, N – f so )a cs B N ⁄ E s ] (4-7) The previous FIP-CEB equation accounts for partial
debonding, but does not consider cyclic creep of the con-
crete. Balaguru (1981) and Balaguru and Shah (1982)
where
developed an equation to estimate maximum crack widths
acs = crack spacing, in. (mm);
that incorporates the effects of cyclic creep and progressive
At = effective area of concrete tension zone surround-
deterioration of flexural stiffness. Comparing the results of
ing all reinforcementhaving the same centroid as
this equation with those of the FIP-CEB crack width formula
the reinforcement, in.2 (mm2);
using data by Bhuvasorakul (1974), Balaguru determined the
P = sum of the perimeters of all tension reinforcement,
following ratios of maximum crack width after N cycles to
in. (mm);
initial maximum crack width: Wmax,N / Wmax = 2.45 (experi-
W = crack width at the first loading cycle; mental) = 1.003 (FIP-CEB) = 2.15 (Balaguru). He attributed
WN = crack width at the Nth loading cycle; the disparity in the result obtained by the FIP-CEB equation
So = slip of the reinforcement at the first cycle; to the lack of consideration of cyclic creep of concrete.
So,N = slip of the reinforcement at the Nth cycle;
B = distance ratio at the first cycle; 4.6—Summary of serviceability
BN = distance ratio at the Nth cycle; In all types of concrete members, crack widths and
fso = steel stress at the primary crack at the first cycle; deflections generally increase under repeated loadings
fso,N = steel stress at the primary crack at cycle N; (Naaman 1982b). These increases occur during early load
N = number of load cycles; and stages and then generally stabilize until approximately 90%
Es = modulus of elasticity of steel. of their fatigue life. Researchers (Harajli and Naaman 1989,
423.5R-20 ACI COMMITTEE REPORT

Fig. 4.6—Comparison of observed and predicted crack width increases with number of
loading cycles (Harajli and Naaman 1989); 1 in. = 25.4 mm.

Shahawi and Batchelor 1986) show a nearly linear relationship cyclic creep of the concrete and tension stiffening effects to
between maximum crack width and reinforcement stress. estimate maximum crack widths and deflections at a given
Deflections can increase to twice the static deflection of the first load cycle. Naaman and Founas (1991) also presented
load cycle. models to compute stresses, strains, and deflections in
The importance of correct reinforcement detailing for partially prestressed beams for static and cyclic loading
crack control should be emphasized. Pretensioned wires or incorporating effects of shrinkage, creep, and relaxation.
nonprestressed bars can be placed near the tensile face to
help limit crack widths according to the report Partial Pre- CHAPTER 5—EFFECTS OF LOAD REVERSALS
stressing (Concrete Society 1983). The presence of ordinary 5.1—Introduction
reinforcement in prestressed beams helps to control cracking Cyclic load reversals can result from dynamic effects of
in static and fatigue tests. Ordinary reinforcement reduces earthquake, shock, wind, or blast loadings. Earthquake load-
the increased steel stresses attributed to cyclic creep of con- ings are imparted to the base of the structure through ground
crete. In addition, it minimizes bond redistribution. motion, whereas wind loadings transmit forces to the above-
Balaguru (1981) and Balaguru and Shah (1982) have ground portions of the structure. Shock or blast loadings can
derived a refined approach that incorporates the effects of involve both of these effects: air overpressure forces similar
PARTIALLY PRESTRESSED CONCRETE 423.5R-21

Fig. 4.7—Crack width variation versus reinforcing steel stress at different PPR (Harajli
and Naaman 1985); 1 ksi = 6.9 MPa, 1 in. = 25.4 mm.

to wind loadings and ground motions accompanying either to be designed for considerably lower forces than required
buried, air, or contact blasts. for the first option.
Designing partially prestressed concrete structures in accor-
5.2—Design philosophy for seismic loadings dance with the second approach, raises the same concerns as
In design for seismic effects, the primary concern is safety for the case of designing with conventional reinforced con-
of the occupants (Bertero 1986), and a secondary concern is crete; that is, the designer must ensure:
economics. Even for an earthquake with reasonable proba- • Ductility at plastic hinges; and
• Adequate energy dissipation through damping and
bility of occurrence; however, it is not possible to guarantee
hysteresis.
with absolute certainty complete life safety and no structural This chapter concentrates on these characteristics as they
damage. The Federal Emergency Management Agency, in relate to partially prestressed concrete.
their document NEHRP 1997 Recommended Provisions
(1998), give minimum requirements to provide “reasonable 5.3—Ductility
and prudent life safety.” To achieve these objectives, struc- Ductility allows the redistribution of forces in indeterminate
tures should be proportioned to resist design forces with ade- systems and ensures gradual rather than brittle failure, provid-
quate strength and stiffness to limit damage to acceptable ing a warning to the occupants before collapse. If adequate
levels. Design forces are determined by using a response- ductility is not provided at the critical regions (plastic hinge
modification coefficient R to reduce the forces obtained, zones), the member will be unable to develop the required
assuming that the structure responds linear elastically to the inelastic rotation.
ground motion. The reduction factor accounts for the Ductility is usually expressed for structural members and
systems in terms of a deformation ductility ratio, where the
reduced strength demand due to a number of effects, includ-
deformation is described in terms of displacement, rotation,
ing energy dissipation through hysteretic damping and
or curvature (Naaman et al. 1986, Thompson and Park
lengthening of the structural period as the structure becomes
1980a, Giannini et al. 1986). Deflection and rotation ductili-
inelastic. ties give an indication of drift (ratio of lateral displacement
In simplistic terms, there are two basic options in seismic to height) and are used in structural analysis. Curvature duc-
design: (1) make the structure strong enough so that it will tility is used to define member or section behavior at plastic
respond elastically; or (2) permit the structure to deform hinges.
inelastically while ensuring adequate ductility and energy The ductility ratio is defined as the ratio of the ultimate
dissipation capacity. The second option permits the structure deformation to the yield deformation. In the case of partially
423.5R-22 ACI COMMITTEE REPORT

frames indicated that the ratios of section curvature ductili-


ties to deflection ductilities of the frame were 1.24 and 1.53
for partially prestressed concrete systems with bonded and
unbonded tendons, respectively.
5.3.1 Factors affecting ductility—Numerous investiga-
tions have been conducted to determine the effect of differ-
ent parameters on ductility. Generally, factors that affect the
ductility of ordinary-reinforced concrete sections affect the
ductility of partially prestressed concrete sections as well.
Effects of anchorage, bond, transfer lengths, grouting, char-
acteristics of high-strength prestressing steels, and level of
prestressing are additional parameters uniquely important to
prestressed and partially prestressed concrete members
(Bertero 1986).
Investigations were conducted to determine the effects of
the reinforcement ratio (Cohn and Ghosh 1972; MacGregor
1974; Scott et al. 1982; and Park and Thompson 1977),
concrete confinement (Park and Falconer 1983; Scott et al.
1982; Burns 1979; Kent and Park 1966; Wight and Sozen
1975; Sheikh and Uzumeri 1982; and Sheikh 1982), material
strengths and stress-strain properties (Wight and Sozen 1975,
Wang 1977), section geometry, ratio of compression
reinforcement (Park and Thompson 1977, Burns and Seiss
1962), PPR, and axial load (Lin and Burns 1981). Another
Fig. 5.1—Definitions of yield for determination of ductility
(Naaman et al. 1986). investigation by Thompson and Park (1980a) was conducted
to determine the effects of the content and distribution of
prestressed concrete structures, the definition of yield longitudinal reinforcement, transverse reinforcement, and
deformation can be quite arbitrary (Naaman et al. 1986; concrete cover. Naaman et al. (1986) analyzed beams with
Thompson and Park 1980a; Giannini et al. 1986). The four types of sections (rectangular, T, I, and box sections),
section contains both prestressed and nonprestressed five different concrete strengths, three types of prestressing
reinforcement that can have different yield stresses; steel, four levels of PPR, and seven reinforcing indexes ω.
prestressing steel does not have a definite yield point, and For the rectangular and T-sections, five levels of effective
reinforcement located in different layers will yield at prestress fpe /fpu were analyzed, and in selected cases, four
different member deformations. The yield deformation, levels of compression reinforcement ratios and the effect of
therefore, can be defined in many ways. confinement were investigated. The following observations
Among numerous investigations on partially prestressed were made from the parametric evaluation:
concrete, there has not been a consistent definition of yield • Reinforcement index—Section ductility decreases with
deformation (Fig. 5.1). Cohn and Bartlett’s study (1982) on increasing ω because of the significant reduction in
partially prestressed concrete assumed that the yield curva- ultimate curvatures associated with increased
ture corresponded to yielding of ordinary reinforcement. reinforcement ratios (Fig. 5.2 and 5.3).
• Effective prestress—When the effective prestress was
This approach gives a relatively high value of ductility
decreased, it led to an increase in ultimate curvature for
because the prestressing reinforcement yields at a much larg- reinforcing indices ω greater than approximately 0.12.
er deformation. Thompson and Park (1980a) defined first It also led to an increase in yield curvature irrespective
yield as the intersection of the tangent to the elastic portion of the reinforcement index. Fig. 5.2 illustrates the
of the load-deformation curve and a horizontal line at ulti- influence of effective prestress on ductility factor or the
mate load. In Park and Falconer’s study (1983) on pre- ratio of these two quantities. The increase in yield
stressed piles, the yield deformation was taken at the curvature tended to dominate the results and caused a
intersection of the secant from zero to 75% of the ultimate reduction in ductility with decreased effective prestress
moment capacity and the postelastic slope. Naaman et al. irrespective of the reinforcing indices. This effect was
(1986) defined the yield deformation as the intersection of more significant for fully prestressed beams.
the secant from zero to the proportional limit of the prestress- • Partially prestressed ratio—A decrease in PPR led to
an increase in both ultimate and yield curvatures, but in
ing steel and the postelastic slope. Although these latter def-
terms of ductility, there was no consistent trend. The
initions give slightly different results, they are all based on effect of PPR on ductility became negligible with
the ductility of the member section rather than yielding of a increased concrete compressive strengths.
particular component in the cross section. • Transverse reinforcement—Increases in concrete
Local section ductility demands can be much more confinement resulted in corresponding increases in
severe than overall deflection ductilities of frames. Tests by sectional ductility irrespective of the reinforcement
Muguruma et al. (1980) on partially prestressed concrete index. This effect was also observed in tests by several
PARTIALLY PRESTRESSED CONCRETE 423.5R-23

Fig. 5.2—Typical variation of curvature ductility as function


of effective prestress in steel from analytical study (Naaman
et al. 1986).

other investigators (Watanabe et al. 1980; Muguruma et


al. 1982b; Iyengar et al. 1970; Okamoto 1980; and Fig. 5.3—Comparison of analytical results with prediction
Hawkins 1977). equations (Naaman et al. 1986); γ = ratio of compression-
• Concrete strength—For high values of reinforcement to-tension steel force at ultimate; and 1 ksi = 6.9 MPa.
index (ω > 0.25), the concrete compressive strength did
not have a significant effect on ductility. For low values reinforcement index ω (computed using the web for T-
of ω (< 0.25), however, the use of higher-strength sections).
concrete reduced section ductility as much as 30%, Naaman et al. (1986) found the reinforcement index ω to
particularly in cases of low PPR (ordinary reinforced be an excellent independent variable for describing section
concrete). This trend was also more pronounced for ductility, as it is proportional to c/dctf (c = distance from
rectangular sections as compared with box, T- and I- extreme compression fiber to neutral axis; and dctf = depth
sections. Tests by Muguruma et al. (1983) indicate that
to centroid of tensile steel). Based on the reinforcement ratio
high-strength concrete is effective in improving
ductility, but that the consequences of high-strength and concrete compressive strength, it can be used to describe
tendon fracture and brittle compressive failure of high- all three types of systems: reinforced concrete, prestressed
strength concrete should be considered. concrete, and partially prestressed concrete.
• Section geometry—The section geometry did not have Based on the results of analysis and experimental tests
an appreciable effect on ductility, provided that the on twelve partially prestressed concrete beams (Harajli and
ductility index was expressed as a function of Naaman 1984) described in the previous chapter, Naaman
423.5R-24 ACI COMMITTEE REPORT
where
ω = reinforcement index;
dctf = depth to centroid of tensile force in steel; and
Lp = equivalent plastic hinge length.
In Fig. 5.4, the equations are compared with the results of
several investigations (Kent and Park 1966; Corley 1966;
Mattock 1964; Mattock 1967; Harajli and Naaman 1985c;
Bishara and Brar 1974; and Baker and Amarakone 1964).
The equations show fairly good correlation with the results
(assuming a plastic hinge length Lp of dctf /2), and are recom-
mended by Naaman et al. for use as a first approximation in
the design and detailing of ordinary reinforced, fully pre-
stressed or partially prestressed members.
In a discussion of this work, Loov et al. (1987) recommend
specifying c/d rather than ω to ensure ductility, as is done in
the Standards Association of Australia’s Draft Unified Con-
crete Structure for Buildings (1984) and the CSA’s Design
of Concrete Structures for Buildings. The use of c/d ensures
consistent ductility demands from sections independent of
geometry and concrete compressive strength. To accommo-
date this, Naaman et al. also provided the expressions in
terms of c/dctf, valid for c/dctf between 0.08 and 0.42:

Curvature ductility ratio Plastic rotation (radians)


1
-------------------------------------------
- 1.86 – 0.73c ⁄ d ctf L P
Upper bound 0.73c ⁄ d ctf – 0.053
- --------------
---------------------------------------- (5-4)
621c ⁄ d ctf – 42 d ctf ⁄ 2

Lower 1
-------------------------------------------- 1.86 – 1.28c ⁄ d ctf L P
1.42c ⁄ d ctf – 0.102
- --------------
---------------------------------------- (5-5)
bound 949c ⁄ d ctf – 50 d ctf ⁄ 2
1
----------------------------------------------- 1.88 – 1.15c ⁄ d ctf L P
Average 1.095c ⁄ d ctf – 0.087
----------------------------------------- -------------- (5-6)
766c ⁄ d ctf – 53 d ctf ⁄ 2

where
c = depth to neutral axis from extreme compression
Fig. 5.4—Comparison of prediction equations with fiber;
experimental results (Naaman et al. 1986). dctf = depth to centroid of tensile force in steel; and
Lp = equivalent plastic hinge length.
et al. developed prediction equations for sectional ductility Thompson and Park (1980a) agreed with a form of c/d as
and plastic rotation as a function of the reinforcement index a measure of the ductility in an investigation in which a
ω. Three equations were derived for both the sectional duc- moment-curvature relation was developed applicable to ordi-
tility and plastic rotation to give upper limit, lower limit, and nary reinforced, prestressed, and partially prestressed concrete
average values for ranges of reinforcement index of 0.05 to beams, and verified with experimental tests on symmetrically
0.3 (Fig. 5.3). The upper and lower limits accommodate the reinforced (prestressed and ordinary reinforced) rectangular
effects of parameters described previously, other than rein- beam-column assemblies subjected to cyclic load reversals.
forcement index ω which were shown to have an effect on The model was subsequently used to conduct a parametric
ductility. For example, Naaman et al. suggest the lower- study. Their recommendations are as follows:
bound equation be used to describe cases with high-strength • Effect of prestressing steel content—ACI 318 limits the
concrete, low effective prestress, and high PPR. The upper- prestressing steel content by requiring ωp ≤ 0.36 β1,
where β1 is a parameter used to modify the neutral axis
bound equations are suggested for cases with normal strength
depth c to determine the depth of the equivalent
concrete, high effective prestress, and low PPR. The equations rectangular stress block a. Although the code limits the
are given as follows prestressing steel content, ACI 318 allows this value to
be exceeded if the additional reinforcement is not
Curvature ductility ratio Plastic rotation (radians) considered in the flexural strength calculation. Although
1 1.05 – ω - -------------
LP this can result in sufficient ductility for gravity loading, it
Upper bound ----------------------- ------------------------ - (5-1)
ω – 0.045 850ω – 35 d ctf ⁄ 2 can be insufficient for seismic loading. To ensure
1 LP
adequate ductility, one can limit ωp to less than 0.2
1.05 – 1.65ω- -------------
Lower bound ---------------------------------
1.94ω – 0.086
- ------------------------------
1300ω – 40 d ctf ⁄ 2
- (5-2) (rather than 0.3 for the case of β1 = 0.85). Alternatively,
rather than using the previous equation, one can also
Average
1
----------------------------
- 1.07 – 1.58ω L P
------------------------------- -------------- (5-3) limit a/h ≤ 0.2 (a = depth of rectangular stress block; and
1.5ω – 0.075 1050ω – 45 d ctf ⁄ 2 h = depth of section). This provides a given amount of
PARTIALLY PRESTRESSED CONCRETE 423.5R-25

ductility by specifying the location of the neutral axis order of 1.5 to three. With Park’s suggested limit of 0.2, duc-
independent of tendon positions. This relationship has tilities in excess of four or five can be obtained. Tests by
been adopted by FIP (Cement and Concrete Association Muguruma et al. (1982a) indicate that if the sections do not
of England 1977) and the Standards Association of New contain compression reinforcement, ω < 0.2 can give ductil-
Zealand’s Code of Practice for General Structural ities of only about three. If the reinforcement index is main-
Design and Design Loading for Buildings (1980). tained at less than 0.1, ductilities in excess of 10 can be
• Effect of prestressing steel distribution on ductility—
achieved (Naaman et al. 1986).
Sections with only one tendon have reduced moment
capacity as the concrete in compression deteriorates. Tests by Nakano and Okamoto (1978) on beam-column
Thompson and Park (1980a) recommend the use of at least subassemblages indicated that even when tension reinforce-
two grouted tendons (one near the top and one near the ment is adequate (for example, ω < 0.15 to provide ductility
bottom of the section) for seismic loading, and suggested of five), hoop reinforcement (> 0.5%) is required to provide
locating a third tendon towards the center of the section. the required deflection ductility. Okamoto (1980) carried out
The use of unbonded tendons was generally not recom- similar tests on eleven simply supported beams and nine can-
mended for primary earthquake-resistant members tilever beams and reached similar conclusions.
because of lack of information on the performance of the 5.3.2 Special cases
anchorages (Muguruma 1986). In cases where partially Compression members—Relatively little information is
prestressed beams are designed with nonprestressed rein-
available on partially prestressed concrete columns (Bertero
forcement in the extreme fibers providing at least 80% of
the seismic resistance, however, they indicated that the 1986). Tertea and Onet (1983) analyzed 20 columns with
prestress can be provided by one or more grouted or different PPR ranging from zero to 100%. As observed in the
unbonded tendons in the middle third of the beam depth. beam tests, curvature ductility of columns improved with the
The centrally located prestressing steel is recommended addition of nonprestressed reinforcement. Curvature ductili-
because it can delay cracking and enhance strength with- ty factors of 3.9, 6.1, and 9.8 corresponded with PPR of 100,
out much reduction in ductility. Studies at the National 50, and 0%. Irrespective of the degree of prestressing, col-
Institute of Standards and Technology (NIST) and as part umn ductilities improved with increased confinement.
of the National Science Foundation Precast Seismic Tests by Park and Falconer (1983) on prestressed piles
Structural Systems (PRESSS) program (Cheok and Lew found that the curvature ductility increased with reduced axi-
1990) have explored the use of unbonded prestressing
al load levels and increased confinement by spiral reinforce-
reinforcement through the connections. In the studies at
NIST, central post-tensioning was used in conjunction ment. They discourage the use of hard-drawn reinforcing
with mild steel reinforcement at the top and bottom of wire for spirals as it commenced fracture at displacement
the beam-column interface. The central post-tensioning ductility factors of four to six. They recommend requiring
maintained integrity of the joint, while energy dissipation the wire to have high fracture strains.
was accommodated with the mild steel reinforcement. A Shear—Beams should be proportioned to ensure ductile
study by Priestley and Tao (1993) proposed the use of flexural behavior and to avoid brittle shear behavior (Bertero
lightly stressed, unbonded prestressing steel through the 1986). To accomplish this objective, the FIP Commission
joint (nonlinear-elastic concept). The system offers ductile (Cement and Concrete Association of England 1977) recom-
behavior through crack opening at the interface, while mends assuming material overstrengths on the order of 15%
maintaining a “self-restoring” force. Because the strands in calculating the design shear force from the plastic hinge
do not yield, they work to bring the connection back to its
moments. The New Zealand Code of Practice for General
originally undeformed position. This system has been
studied further with other proposed connection systems Structural Design and Design Loading for Buildings (Stan-
as part of the PRESSS program (Cheok and Lew 1990). dards Association of New Zealand 1980) recommend
• Effect of transverse reinforcement on ductility—The neglecting the concrete contribution to shear resistance when
degree of confinement provided by the transverse reinforce- the axial compression is less than 0.1 f c′ .
ment to the compression regions significantly increases Muguruma et al. (1983) determined that concrete resis-
ductility. The confinement enhances the performance of the tance to shear can be improved by prestressing. Experiments
concrete within the core and also inhibits buckling of the were conducted on 7 x 10 in. (180 x 250 mm) simply sup-
compression reinforcement (Park and Thompson 1977; ported beams with span lengths of 7.22 ft. (2.20 m) to inves-
Park and Paulay 1975; and Thompson 1975). Thompson tigate shear in prestressed beams. The flexural shear
and Park recommend that the stirrups be spaced less than cracking load was found to increase in direct proportion to
d/4 or 6 in. (150 mm).
the increase in flexural cracking load (due to an increase in
• Effect of cover thickness—For ductility, Thompson and
prestressing). Muguruma (1986) suggests in the design of
Park suggest that the cover be made as small as possible,
because loss of cover under cyclic loads causes a reduc- partially prestressed concrete beams that the concrete contri-
tion in capacity. The effect of cover spalling on member bution to shear resistance can be increased by the shear force
strength is not as great for deeper members where cover corresponding to the decompression moment at the critical
is a smaller percentage of member depth. section. This effect is taken into consideration in ACI 318.
Holding all other parameters constant, ductility decreases Beam-column joints—Beam-column joints should be
with an increase in reinforcement index. ACI 318 gives a designed and detailed so that the plastic-hinge zones in the
maximum value for the reinforcement index of ω ≤ 0.36 β1. beams can develop their full plastic capacities. Muguruma
Maintaining this limit would give section ductilities on the (1986) indicates that there are no differences in joint design
423.5R-26 ACI COMMITTEE REPORT

for ordinary reinforced or partially prestressed systems,


except that the prestressing tendons do not bond to the con-
crete as well as ordinary deformed reinforcement. Research
on the performance of bond between tendons and concrete in
joints (Muguruma and Tominaga 1972) has indicated that
tendons undergo bond deterioration at early load stages.
Tendon slip causes a reduction in energy dissipation capacity
by pinching the hysteresis loops.
Park’s investigation of beam-column subassemblages
with grouted post-tensioned tendons (Applied Technology
Council 1981) recommended minimizing the possibility of
failure in the joint. To avoid this type of failure, the hinge
zones should be displaced away from the beam-column
interface. This can be accomplished by haunching the
beams, adding additional reinforcement at the interface, or Fig. 5.5—Moment-curvature idealizations for prestressed,
using cruciform shapes (Bertero 1986). Displacing the hinge reinforced, and partially prestressed concrete (Thompson
from the interface has the added advantage of locating the and Park 1980b); α = ratio of strength of reinforced
concrete component to total strength of partially prestressed
failure in a region in which nonprestressed reinforcement is concrete system; and β = ratio of strength of prestressed
located, increasing ductility and energy dissipation. Displac- concrete component to total strength of partially prestressed
ing the hinge further into the beam results in increased shear concrete system.
at the interface when the hinge reaches its capacity and
produces increased ductility demands relative to hinges at strengths (Thompson and Park 1980b). Prestressed concrete
the interface when subjected to similar drift levels. Using members can achieve significant elastic recoveries even at
post-tensioned tendons through the joint core at middepth large deformations (Schoeder 1977) because of the initial
can reduce the amount of required joint shear reinforce- tensile stress in the strand due to prestressing. In elastic
ment as it serves to maintain integrity of the joint (Park and recoveries, the strand unloads its tensile force and is capable
Thompson 1977).
of achieving large recoveries. The member will not achieve
significant energy dissipation unless the prestressing steel
5.4—Energy dissipation yields, the concrete crushes, or the nonprestressed deformed
Energy dissipation can be attributed to viscous and hyster-
bar yields (Nakano 1965; Blakeley and Park 1973b; Guyon
etic damping described separately below. Energy dissipation
1965; Blakeley and Park 1971; and Bertero 1973). The
of prestressed concrete is as little as 0.15 times that of con-
ventional reinforced concrete of comparable size and addition of nonprestressed reinforcement in partially
strength (Bertero 1986). prestressed concrete members has been shown to provide a
marked improvement in energy dissipation characteristics
5.4.1 Viscous damping—In a 1978 state-of-the-art report
by Hawkins (1977) on seismic resistance of prestressed and (Thompson and Park 1980b, Inomata 1986). This is evident
precast structures, a thorough review of the research on in Fig. 5.5, which shows typical hysteresis loops for
damping was presented. Tests by Penzien (1964) indicated cyclically loaded fully prestressed, ordinary reinforced, and
that the level of prestress and concrete strength affects damp- partially prestressed concrete members. In addition, by
ing by affecting cracking. Decreases in cracking cause a reducing the strength degradation, the presence of
reduction in damping. longitudinal nonprestressed reinforcing steel improves
Tests by Spencer (1969) on centrally prestressed beams ductility by acting as compression reinforcement (Park and
indicated that damping ratios were not frequency dependent. Thompson 1977, Inomata 1986). As mentioned previously,
Damping ratios tended to increase with end rotations and Cheok and Lew (1990) have investigated the use of lightly
prestress. Spencer also found damping to be higher for prestressed unbonded reinforcement. The performance of
beams loaded in uniform shear rather than uniform moment the PRESSS connections has the advantage of a self-
and attributed this result to the higher effects of interface restoring force (nonlinear-elastic system). The stiffness
shear and bond slip in the case of uniform shear loading. reduction compared with that of a monolithic counterpart
Tests by Brondum-Nielsen (1973) on centrally prestressed can compensate for the reduced energy dissipation. Hybrid
beams indicated damping decreased with increased prestress connections (Cheok and Lew 1991) have been investigated
levels, which is the opposite of what Spencer observed. Val- that combine the nonlinear elastic system with elements that
ues of damping obtained by Nakano (1965) from tests of 1/3 yield or dissipate energy.
scale four-story frames were 1% before cracking, 3% at Thompson and Park (1980a, 1980b) developed idealized
cracking, and 7% before inelastic behavior. moment-curvature relations for partially prestressed
5.4.2 Hysteretic behavior—The energy dissipation (area concrete sections by combining relations obtained by
enclosed within the force-deformation hysteresis loops) of Blakeley (1971) and Blakeley and Park (1973a) for
prestressed concrete members is lower than that of prestressed concrete and by Ramberg and Osgood for
reinforced concrete members designed to develop similar reinforced concrete (Iwan 1973). The idealized hysteresis
PARTIALLY PRESTRESSED CONCRETE 423.5R-27

Fig. 5.6—Comparison of idealized and experimental


moment-curvature relations for prestressed concrete section
(Thompson and Park 1980b). Fig. 5.7—Comparison of idealized and experimental
moment-curvature relations for reinforced concrete section
(Thompson and Park 1980b).
curves agree with those obtained experimentally from tests
on beam-column subassemblages. Fig. 5.6 through 5.8 show
the idealized and experimental moment-curvature
relationships for fully prestressed, reinforced, and partially
prestressed members subjected to cyclic loading.
In the fully prestressed member (Fig. 5.6), the energy dissi-
pation capacity of the member increased (although still with
considerable elastic recovery) after the beam-moment capaci-
ty had been reached and crushing of the concrete had com-
menced. The compression steel in the beam was nominal and
the significant degradation in the strength and stiffness was
attributed to the reduction in the sectional area when the cover
concrete crushed.
After the beam-moment capacity had been reached and
crushing of the concrete had commenced for the partially
Fig. 5.8—Comparison of idealized and experimental
prestressed member (Fig. 5.8), the energy dissipation capac- moment-curvature relations for partially prestressed con-
ity of the partially prestressed member was considerably crete section (Thompson and Park 1980b).
greater than that observed for the fully prestressed member.
For this particular test, the beneficial effect of the nonpre- 5.5—Dynamic analyses
stressed longitudinal steel in the compression zone of the Giannini et al. (1986) presented an analytical investigation
beam plastic hinge region was evident from the stable on the seismic behavior of fully prestressed and partially pre-
moment-curvature relations. The strength degradation was stressed concrete. The influences of the fundamental period,
not large because after concrete crushing commenced, the PPR, and seismic intensity were considered. The response
nonprestressed steel was able to carry some of the compres- was obtained in the form of time histories and peak values in
sion previously carried by the cover concrete and therefore terms of displacements, curvatures, local strains (in the con-
stituents), forces, and dissipated energy.
helped maintain the internal lever arm (Thompson and Park
1980b). Partially prestressed concrete seems to couple the
advantages of fully prestressed concrete with those of ordinary
Similar findings were obtained by Muguruma et al.
reinforced concrete in resisting loads. Fully prestressed
(1982a, 1982b), who tested 24 partially prestressed (eccen-
concrete has a tendency to greater strain and deformation
tric) beams with various levels of prestress, and also by
demands than ordinary reinforced concrete. After a strong
Nakano and Okamoto (1978), who tested 21 partially pre- impulse, the displacements of the reinforced concrete and
stressed (symmetric) beams subjected to cyclic-load reversals. partially prestressed concrete systems tended to damp out,
Confining the concrete has a marked benefit on the ductility whereas those of the fully prestressed concrete system did
of partially prestressed concrete and also improves the hyster- not. The analysis showed that fully prestressed concrete
etic behavior. Closely spaced stirrup ties or spirals in the hing- resists the design seismic intensity, but for greater
ing regions enhance ductility and energy dissipation by intensities, it rapidly approaches collapse. Fully prestressed
confining the concrete compression zone, reducing stiffness concrete structures, however, can behave better for low
degradation, and inhibiting buckling of the compression steel. intensity seismic excitation.
423.5R-28 ACI COMMITTEE REPORT

Similar conclusions were obtained from a study by connections. The most viable alternatives were continuous
Thompson and Park (1980b). They used the idealized post-tensioned connections displacing the plastic-hinge
moment-curvature relations (discussed previously) to con- region from the beam-column interface, and tapered-
duct a nonlinear dynamic analysis of ordinary reinforced, threaded splice connections that emulate the performance of
fully prestressed, and partially prestressed concrete single- monolithically cast beam-column connections. It was also
degree-of-freedom systems. Displacements of fully pre- suggested that advantage be taken of composite action
stressed concrete systems are larger than those of ordinary between the beam and the slab.
reinforced and partially prestressed concrete systems
because of the reduced energy dissipation associated with
5.7—Summary
prestressing.
If members are detailed properly, it is possible to ensure
Fully prestressed systems also undergo a significant
ductile behavior in fully prestressed and partially prestressed
reduction in stiffness after cracking that counteracts the
concrete beams (Applied Technology Council 1981).
increased displacements under dynamic loading. The reduc-
tion in stiffness causes a reduction in the effective period of Ordinary reinforced concrete has an advantage in energy-
vibration that can reduce the displacement response. Under dissipation characteristics. Prestressed concrete has an
the combination of both of these effects, the analyses indi- advantage in its ability to reach large displacements with
cate that fully prestressed frames will have, on average, 30% only a small loss in load-carrying capacity and almost com-
greater maximum lateral displacements than reinforced con- plete recovery of displacements. Joint performance is
crete systems designed to have the same strength, viscous improved by the use of continuous prestressed strand
damping ratio, and initial stiffness (Thompson and Park through the joint.
1980b). Partially prestressed concrete couples the advantages of
Similar analytical studies have been conducted as part of prestressed concrete with those of ordinary reinforced con-
the NSF PRESSS program (Palmieri et al. 1996, El Sheikh crete (Giannini et al 1986). The sections can be detailed so
et al. 1999) that indicated systems incorporating unbonded that energy dissipation and deformation demands are on the
post-tensioned connections experienced moderate increases same order as those of ordinary reinforced concrete. Partially
in displacements in comparison with those expected from prestressed concrete also offers the possibility of calibrating
monolithic concrete frame systems. An advantage of the
the subsequent yielding of both prestressed and nonpre-
unbonded post-tensioned connections was the reduction in
stressed of reinforcement to optimize stiffness and energy-
residual drift associated with the nonlinear elastic behavior
dissipation properties (Giannini et al 1986).
of the system in comparison with the permanent system
deformations associated with the inelastic response of the A recommended design concept is to balance all or part of
ordinary reinforced frames. the gravity loads, satisfy serviceability criteria with draped
Fully prestressed concrete members can be detailed to or blanketed prestressing strands, and place sufficient non-
accommodate increased displacements in a ductile manner; prestressed reinforcement in the section to meet the addition-
however, these higher displacements can result in greater al requirements for seismic design strength, ductility, and
nonstructural damage and P-delta effects. An advantage of energy dissipation (Hawkins 1977; Inomata 1986; and
using full prestressing and partial prestressing is the reduc- Thompson and Park 1980b). Stirrup ties should be located in
tion in residual displacements due to the self-restoring nature hinging regions to prevent bar buckling, confine concrete,
of these systems. The addition of nonprestressed reinforce- and increase shear strength (Park and Thompson 1977, Park
ment can reduce these deflections. and Paulay 1975).

5.6—Connections CHAPTER 6—APPLICATIONS


Results of tests (Berg and Degenkolb 1973; Elliott 1972; 6.1— Early applications
Kunze et al. 1965; Sutherland 1965; Ozaka 1968; FIP 1970; Partial prestressing has been used in the design of
FIP 1974; Arya 1970; and Architectural Institute of Japan
structures since the early 1950s. In 1952, partially prestressed
1970) and observations on the past performance of pre-
beams with post-tensioned cables were first used in the roof
stressed concrete in earthquakes indicate that most pre-
of a freight depot at Bury St. Edmunds, England (Bennet
stressed concrete members perform well. Most failures have
1984). Lower concrete tensile stresses were specified
been found to occur in connections between elements.
initially, but with successful experience and an increasing
Several investigations have been conducted to study the
performance of connections between fully prestressed and amount of test data, the permissible fictitious tensile stresses
partially prestressed concrete precast elements. As mentioned were increased to 750 psi (5.2 N/mm2) for pretensioning and
previously, ductile connection concepts have been tested as 650 psi (4.5 N/mm2) for post-tensioning (Bennet 1984).
part of the PRESSS program. French et al. (1989a, 1989b), Other early applications of partial prestressing included the
Amu and French (1985), Hafner and French (1986), development of a concrete mast for overhead power lines
Jayashankar and French (1987), and Tarzikhan and French and structures designed to withstand mining collapse; with
(1987) investigated post-tensioned, welded, bolted, both applications, the most severe design load was usually a
composite, threaded coupler, and tapered-threaded splice temporarily unbalanced load (Bennet 1984).
PARTIALLY PRESTRESSED CONCRETE 423.5R-29

6.2—Pretensioned concrete components


In the early 1980s, a survey was conducted among the pre-
tensioned, precast-concrete plants in the U.S. and Canada
(Jenny 1985) concerning the use of partial prestressing in the
industry. Sixty-five plants responded to the survey. Four of
the plants manufactured only bridge girders, and partial pre-
stressing was rarely considered in their production. The sur-
vey revealed, however, that there was a general familiarity
with partial prestressing and the concept was used for meet-
ing special design situations rather than as a routine produc-
tion standard. Only seven of the 61 remaining plants
responding to the survey indicated that they did not allow
any product to exceed a nominal tensile stress of 6√f c′ psi
(√f c′ /2 MPa) for certain members, such as stemmed units,
inverted tees, spandrel beams, hollow-core slabs, or columns
(Jenny 1985).
The nominal stress levels reported by the plants were: 7.5
to 12√f ′c psi (√f ′c / 1.6 to √f ′c MPa) for stemmed units; 7.5
to 12√f ′c psi (√f ′c /1.6 to √f ′c MPa) for beams; 7.5 to 9√f ′c
psi (√f ′c / 1.6 to √f c′ /1.33 MPa) for hollow-core slabs; and
7.5√f c′ psi (√f c′ /1.6 MPa) for columns. Reasons for design- Fig. 6.1—Structural floor plan layout of 1100 Peachtree
ing these members with a nominal tensile stress greater than Office Tower, Atlanta, Ga.
6√f c′ psi (√f c′ /2 MPa) were to meet special design loads,
reduce camber, save steel and lower the cost, reduce the of tendons and by reducing the overall height of the structure,
required concrete strength at release, and achieve competi- owing to the shallower beam depths.
tive span-depth ratios (Jenny 1985). For two-way flat-plate designs, ACI 318 limits the stress
at column locations as calculated by the equivalent frame
6.3—Post-tensioned building construction method or other approximate methods to 6√f c′ ( √f c′ /2
The majority of post-tensioned building construction in MPa). This is in recognition of the gross approximation of
the U.S. is based on the use of unbonded tendons in conjunc- the actual strains in the vicinity of the column. Higher allow-
tion with code-specified minimum amounts of bonded non- able stresses, however, are permitted in conjunction with
prestressed reinforcement. Maximum design tensile stresses more precise analytical procedures.
range from 6 to 12√f c′ psi (√f c′ /2 to √f c′ MPa) (Feyermuth
1985). For designs where durability considerations do not 6.4—Bridges
govern, service-load nominal tensile stress levels of 6 to Prestressed concrete bridges designed to permit cracking
12√f c′ psi (√f c′ /2 to √f c′ MPa) are now routine for one-way, under service loads are still rare in the U.S., even though it is
post-tensioned systems. According to Feyermuth, “The use common to allow flexural tension in extreme fibers under
of partial prestressing, especially as applied to flat slabs, has service load. The primary concerns about bridge members
reduced the incidence of significant cracking problems in being cracked under service loads have to do with fatigue
post-tensioning buildings. This results from lower prestress and durability. Partially prestressed concrete bridges have
levels, inducing proportionately smaller elastic and creep been built in Europe and Japan.
shortening movements, and the presence of bonded nonpre- Taerwe (1990) has described a number of partially pre-
stressed reinforcement to control crack widths” (1985). stressed concrete bridges constructed in Belgium. These
In the design of office and hotel structures, the use of partial highway bridges were designed according to the Class 2 con-
prestressing is often cost-effective. Post-tensioned beams, as ditions described in Chapter 1 of this report. As of 1990, the
well as one-way slabs, designed to an extreme nominal fiber bridges described by Taerwe (built between 1965 and 1975)
stress in tension between 6 and 12√f c′ psi (√f c′ /2 and √f c′ performed satisfactorily.
MPa) have performed well in practice. The use of partially The Kannonji Viaduct in Japan was completed in 1991
prestressed design is especially effective when certain (Soda and Kazuo 1994). The bridge has a total length of
designated areas of a building are required to have a higher 2464 ft (751 m), with an average span length of 92 ft (28 m).
design live load than the majority of the floor area. The 1100 The bridge superstructure units are continuous for three or
Peachtree Tower (shown in Fig. 6.1) is a 27-story, cast-in- four spans each. It was designed to allow flexural cracking
place concrete office building located in Atlanta, Ga. The slab under service loads at midspan but not over interior supports.
is 5 in. (127 mm) thick, conventionally reinforced, and Four other partially prestressed concrete highway bridges
supported by post-tensioned beams 14 in. (356 mm) wide and also have been constructed in Japan (Soda and Kazuo 1994).
22 in. (559 mm) deep. By allowing the nominal tensile stress AASHTO’s Standard Specifications for Highway Bridges
in the beams to reach 10√f c′ psi (√f c′ /1.2 MPa) under service does not currently allow service load cracking in prestressed
loads, economy was achieved by reducing the overall number concrete bridges. AASHTO, however, has adopted a load-
423.5R-30 ACI COMMITTEE REPORT

and resistance-factor design specification for highway bridg- The reports and standards listed above were the latest edi-
es (AASHTO LRFD Bridge Design Specifications) that does tions at the time this document was prepared. Because these
specifically address partially prestressed concrete. This can documents are revised frequently, the reader is advised to
lead to a wider use of partially prestressed concrete for high- contact the proper sponsoring group if it is desired to refer to
way bridges in the U.S. the latest version.

6.5—Other applications AASHTO


Dolan* reports that the prismatic beams on the Walt Dis- 444 N. Capitol Street NW
ney World Monorail were designed for tensile stresses of Suite 249
nearly 12√f c′ psi (√f c′ MPa) for certain load combinations. Washington D.C. 20001
The beams were precast and post-tensioned to provide con-
tinuity. Reinforcing bars were used to control cracking at the American Concrete Institute
continuity joint. P.O. Box 9094
Partial prestressing also has applications in nuclear-con- Farmington Hills, Mich. 48333-9094
tainment structures, cylindrical shells, and rectangular water
vessels. In the 1960s, the first prestressed concrete contain- ASTM
ment structures for nuclear reactors were partially pre- 100 Barr Harbor Drive
stressed in the vertical direction only with nonprestressed West Conshohocken, PA 19428
reinforcement in the circumferential (hoop) direction of the
cylinder and in the dome. Two examples are the containment 7.2—Cited references
structures at R.E. Ginna and H.B. Robinson Unit 2 nuclear Abeles, P. W., 1940, “Saving Reinforcement by Prestress-
stations (Ashar et al. 1994). Similarly, partial prestressing is ing,” Concrete and Constructional Engineering, V. 35, 3 pp.
used vertically in the walls of cylindrical shell and rectangu- Abeles, P. W., 1945, “Fully and Partially Prestressed
lar water-storage structures. If these structures are designed Reinforced Concrete,” ACI JOURNAL, Proceedings V. 41,
with restrained bases, high negative or positive vertical No. 3, Jan., pp. 181-214.
bending moments, or both, will be present. An economical Abeles, P. W., 1954, “Static and Fatigue Tests on Partially
Prestressed Concrete Constructions,” ACI JOURNAL, V. 51,
design can be produced when a small amount of vertical pre-
No. 12, Dec., 361 pp.
stressing in the range of 200 psi (1.4 MPa) is provided and
Abeles, P. W., 1965, “Studies of Crack Widths and Defor-
nonprestressed reinforcement is added to resist any net flex- mation Under Sustained and Fatigue Loading,” PCI Journal,
ural tensile stresses. These high bending moments occur V. 10, No. 6, Dec., pp. 43-52.
only at one point in the wall. Therefore, the economy is Abeles, P. W., 1968, “Preliminary Report on Static and
achieved by not overreinforcing the entire height of wall. Sustained Loading Tests,” PCI Journal, V. 13, No. 4, Aug.,
This method of design has been used since the 1970s.† pp. 12-32.
ACI Committee 209, 1982, “Prediction of Creep, Shrink-
CHAPTER 7—REFERENCES age, and Temperature Effects in Concrete Structures,”
7.1—Referenced standards and reports Designing for Creep and Shrinkage, SP-76, B. B. Goyal,
American Concrete Institute ed., American Concrete Institute, Farmington Hills,
Mich., pp. 193-300.
222R Corrosion of Metals in Concrete
ACI Committee 215, 1974, “Considerations for Design of
318 Building Code Requirements for Structural Con- Concrete Structures Subjected to Fatigue Loading,” ACI
crete and Commentary JOURNAL, V. 71, No. 3, Mar., pp. 97-121.
343R Analysis and Design of Reinforced Concrete ACI Committee 224, 1980, “Control of Cracking in Con-
Bridge Structures crete Structures,” Concrete International, V. 2, No. 10, Oct.,
423.3R Recommendations for Concrete Members Pre- pp. 35-76.
stressed with Unbonded Tendons ACI-ASCE Joint Committee 323, 1958, “Tentative Rec-
ommendations for Prestressed Concrete,” ACI JOURNAL,
ASTM Proceedings V. 54, No. 7, Jan., pp. 545-578.
A 416 Standard Specification for Uncoated Seven- Al-Zaid, R. Z., and Naaman, A. E., 1986, “Analysis of
Wire Stress-Relieved Strand for Prestressed Partially Prestressed Composite Beams,” Journal of Struc-
Concrete tural Engineering, ASCE, V. 112, No. 4, Apr., pp. 709-725.
American Association of State Highway and Transportation Al-Zaid, R. Z.; Naaman, A. E.; and Nowak, A. S., 1988,
Officials (AASHTO) “Analysis of Partially Prestressed Composite Beams under
AASHTO/ Bridge Design Specifications Sustained and Cyclic Fatigue Loading,” Journal of Struc-
tural Engineering, ASCE, V. 114, No. 2, Feb., pp. 269-291.
LRFD Amu, O. O., and French, C. W., 1985, “Moment Resistant
AASHTO Standard Specification for Highway Bridges Connections in Precast Structures Subjected to Cyclic Lat-
eral Loads,” Structural Research Report No. 87-07,
Deptartment of Civil and Mineral Engineering, University
* April 16, 1991 corresponence from Charles W. Dolan to Paul Zia.
of Minnesota, Minneapolis, Minn.
† October 26, 1995 correspondence from Steven R. Close to Nick Marianos. Applied Technology Council, 1981, Proceedings, Work-
PARTIALLY PRESTRESSED CONCRETE 423.5R-31

shop on Design of Prefabricated Concrete Buildings for ity of Prestressed Concrete Beams,” Journal of the Struc-
Earthquake Loads, Los Angeles, Calif., Apr., tural Division, ASCE, V. 100, No. ST9, Sept., pp. 1883-
Architectural Institute of Japan, 1970, Design Essentials 1895.
in Earthquake Resistant Buildings, Elsevier Co., New York, Blakeley, R. W. G., 1971, “Ductility of Prestressed Con-
pp. 186-200. crete Frames under Seismic Loading,” PhD thesis, Univer-
Arya, A. S., 1970, “Use of Prestressed Concrete in Earth- sity of Canterbury, Christchurch, New Zealand.
quake Resistance Structures,” Proceedings, Fourth Sympo- Blakeley, R. W. G., and Park, R., 1971, “Seismic Resis-
sium on Earthquake Engineering, Roorkee, India, Nov. tance of Prestressed Concrete Beam-Column Assemblies,”
Ashar, H.; Naus, D.; and Tan, C.P., 1994, “Prestressed ACI JOURNAL, V. 68, No. 9, Sept., pp. 677-692.
Concrete in U.S. Nuclear Power Plants (Part 1),” Concrete Blakeley, R. W. G., and Park, R., 1973a, “Response of
International V. 16, No. 5, May, pp. 30-34. Prestressed Concrete Structures to Earthquake Motions,”
Baker, A. L. L., and Amarakone, A. M. N., 1964,, “Inelas- New Zealand Engineering, V. 28, No. 2, Feb., pp. 42-54.
tic Hyperstatic Frame Analysis,” Proceedings, International Blakeley, R. W. G., and Park, R., 1973b, “Prestressed Con-
Symposium on the Flexural Mechanics of Reinforced Con- crete Sections With Seismic Loadings,” Journal of the Struc-
crete, SP-12, American Concrete Institute, Farmington tural Division, ASCE, V. 99, No. ST8, Aug., pp. 1717-1742.
Hills, Mich., pp. 85-142. Branson, D. E., 1974, “The Deformation of Non-Compos-
Balaguru, P. N., 1981, “Analysis of Prestressed Concrete ite and Composite Prestressed Concrete Members,” Deflec-
Beams for Fatigue Loading,” PCI Journal, V. 26, No. 3, tions of Concrete Structures, SP-43, American Concrete
May-July, pp. 70-94. Institute, Farmington Hills, Mich., pp. 83-128.
Balaguru, P. N., and Shah, S. P., 1982, “Method of Pre- Branson, D. E., 1977, Deformation of Concrete Struc-
dicting Crack Widths and Deflections for Fatigue Loading,” tures, McGraw-Hill, New York.
Fatigue of Concrete Structures, SP-75, S. P. Shah, ed., Branson, D. E., and Kripanarayanan, K. M., 1971, “Loss
American Concrete Institute, Farmington Hills, Mich., pp. of Prestress, Camber, and Deflections of Non-Composite
153-175. and Composite Prestressed Concrete Structures,” PCI Jour-
Beeby, A. W., 1978, “Corrosion of Reinforcing Steel in nal, V. 16, No. 5, Sept.-Oct. pp. 22-52.
Concrete and its Relation to Cracking,” Structural Engi- Branson, D. E., and Shaikh, A. F., 1985, “Deflection of
neer, London, V. 56A, No. 3, Mar., pp. 77-81. Partially Prestressed Members,” Deflections of Concrete
Beeby, A. W., 1979, “The Prediction of Crack Widths in Structures, SP-86, G. M. Sabnis, Ed., American Concrete
Hardened Concrete,” Structural Engineer, London, V. 57A, Institute, Farmington Hills, Mich., pp. 323-363.
No. 1, Jan., pp. 9-17. Branson, D. E., and Trost, H., 1982a “Unified Procedures
Bennet, E. W., 1984, “Partial Prestressing—An Historical for Predicting the Deflection and Centroidal Axis Location
Overview,” PCI Journal, V. 29, No. 5, Sept.-Oct., pp. 104-117. of Partially Cracked Nonprestressed and Prestressed Con-
Bennett, E. W., 1986, “Partially Prestressed Concrete Mem- crete Members,” ACI JOURNAL, V. 79, No. 2, Mar.-Apr., pp.
bers: Repeated Loading,” Partial Prestressing, From Theory 119-130.
to Practice, NATO ASI Series, V. 1, Martinus Nijhoff Pub- Branson, D. E., and Trost, H., 1982b “Application of the I-
lishers, Dordrecht, The Netherlands, pp. 135-149. Effective Method in Calculating Deflections of Partially Pre-
Bennett, E. W., and Lee, K. H., 1985, “Deflection of Par- stressed Members,” PCI Journal, V. 27, No. 5, Sept.-Oct.
tially Prestressed Beams under a Combination of Long- Brondum-Nielsen, T., 1973, “Effect of Prestress on the
Time and Short-Time Loading”, Deflections of Concrete Damping of Concrete,” Resistance and Ultimate Deform-
Structures, SP-86, G. M. Sabnis, Ed., American Concrete ability of Structures Acted on by Well Defined Repeated
Institute, Farmington Hills, Mich., pp. 185-214. Loads, IABSE, Lisbon.
Berg, G. V., and Degenkolb, H. J., 1973, “Engineering Les- Bruggeling, A. S. G., 1977, “Time Dependent Deflection
sons from the Managua Earthquake,” Managua, Nicaragua of Partially Prestressed Beams,” Report No. 5-77-1, Depart-
Earthquake, Dec. 1972, Earthquake Engineering Research ment of Civil Engineering, Delft Technical University, Mar.,
Institute, Oakland, Calif., Nov., pp. 761-763. 16 pp.
Bertero, V. V., 1973, “Experimental Studies Concerning Bureau of Public Roads, 1954, Criteria for Prestressed
Reinforced, Prestressed, and Partially Prestressed Concrete Concrete Bridges, U.S. Department of Commerce, Wash-
Structures and their Elements,” Resistance and Ultimate ington, D.C.
Deformability of Structures Acted on by Well Defined Burns, N. H., 1964, “Moment Curvature Relationship for
Repeated Loads, IABSE, Lisbon. Partially Prestressed Concrete Beams,” PCI Journal, V. 9,
Bertero, V. V., 1986, “Partially Prestressed Concrete No. 1, Feb., pp. 52-63.
Members for Earthquake-Resistant Design and Construc- Burns, N. H., and Siess, C. P., 1962, “Loading Deforma-
tion,” Partial Prestressing, from Theory to Practice, NATO tion Characteristics of Beam-Column Connections in Rein-
ASI Series, V. 1, Martinus Nijhoff Publishers, Dordrecht, forced Concrete,” Civil Engineering Studies, Structural
The Netherlands, pp. 151-188. Research Series, Bulletin No. 234, University of Illinois,
Bhuvasorakul, T., 1974, “Performance of Members Rein- Urbana, 261 pp.
forced with Smooth Welded Wire Fabric Under Static and Burns, V. V., 1979, “Inelastic Behavior of Structural Ele-
Repeated Loads,” PhD thesis, Oklahoma State University. ments and Structures,” Proceedings, NSF Workshop (Dec.
Birkenmaier, M., 1984, “Developments in the Design of 2-4), University of Illinois at Chicago, pp. 96-167.
Partial Prestressing,” Partial Prestressing, from Theory to Caflisch, R.; Krauss, R.; and Thurlimann, B., 1971,
Practice, NATO ASI Series, Martinus Nijhoff Publishers, “Beige-und Schubversuche and Teilweise Vorgespannten
Dordrecht, The Netherlands. Betonbalken, Serie C.,” Bericht 6504-3, Institut fur Bausta-
Bishara, A. G., and Brar, G. S., 1974, “Rotational Capac- tik, ETH, Zurich.
423.5R-32 ACI COMMITTEE REPORT

CEB-FIP; 1978, Model Code for Concrete Structures, “Report on Prestressing Steels: Types and Properties,” Fed-
Paris, 348 pp. eration Internationale de la Precontrainte (FIP), London,
Cement and Concrete Association, 1977, Recommenda- England, Aug.,18 pp.
tions for the Design of Aseismic Prestressed Concrete Struc- FIP Commission on Prestressing Steel and Systems, 1992,
tures, London, Nov., 28 pp. “Recommendations for Acceptance of Post-Tensioning Sys-
Cheok, G. S., and Lew, H. S., 1990, “Performance of 1/3- tems (Draft Version),” Federation Internationale de la Pre-
Scale Model Precast Concrete Beam-Column Connections contrainte (FIP), London, England, March 10, 21 pp.
Subjected to Cyclic Inelastic Loads,” NISTIR 4433, NIST, FIP Commission on Seismic Structures, 1970, “Report of
Gaithersburg, Md. the FIP Commission on Seismic Structures,” Proceedings,
Cheok, G. S., and Lew, H. S., 1991, “Performance of 1/3- Sixth Congress of FIP (Prague), Federation Internationale
Scale Model Precast Concrete Beam-Column Connections de la Precontrainte (FIP), London, England, pp. 91-95.
Subjected to Cyclic Inelastic Loads — Report No. 2,” NIS- FIP Commission on Seismic Structures, 1974, “Report of
TIR 4589, NIST, Gaithersburg, Md. the FIP Commission on Seismic Structures,” Proceedings,
Cohn, M. Z., and Bartlett, M., 1982, “Nonlinear Flexural Seventh Congress of FIP (New York), Federation Internatio-
Response of Partially Prestressed Concrete Sections,” Jour- nale de la Precontrainte (FIP), London, England, pp. 64-73.
nal of the Structural Division, ASCE, V. 108, No. ST12, FIP-CEB Joint International Committee, 1970, “Recom-
Dec., pp. 2747-2765. mendations for the Design and Construction of Concrete
Cohn, M. Z., and Ghosh, S. K., 1972, “The Flexural Duc- Structure,” Cement and Concrete Association, London,
tility of Reinforced Concrete Sections,” Publications, Inter- England, June, 80 pp.
national Association for Bridge and Structural Engineering, Founas, M, 1989, “Deformation and Deflections of Partially
Zurich, V. 32-II, pp. 53-83. Prestressed Concrete T-Beams under Static and Random
Corley, W. G., 1966, “Rotational Capacity of Reinforced Amplitude Fatigue Loading,” PhD thesis, Department of Civil
Concrete Beams,” Journal of the Structural Division, Engineering, University of Michigan, 402 pp.
ASCE, V. 92, No. ST5, Oct., pp. 121-146. (See also PCA French, C. W.; Amu, O.; and Tarzikhan, C., 1989a, “Con-
Bulletin D108.) nections between Precast Elements—Failure Outside Con-
CSA Committee A23.3, 1984, Design of Concrete Struc- nection Region,” Journal of Structural Engineering, ASCE,
tures for Buildings, CAN3-A23.3-M84, Canadian Standards V. 115, No. 2, Feb., pp. 316-340.
Association (CSA), Rexdale, Ontario, Canada, 281 pp. French, C. W.; Hafner, M.; and Jayashankar, V., 1989b,
de Saint-Venant, 1956, “Memoires presentes par divers “Connections between Precast Elements—Failure Within
savants a I’Academie des Sciences de I’Institut Imperial de Connection Region,” Journal of Structural Engineering,
France,” V. 14 (Second Series), Paris. ASCE, V. 115, No. 12, Dec., pp. 3171-3192.
Dilger, W. H., 1982, “Creep Analysis of Prestressed Con- Freyermuth, C. L., 1985, “Practice of Partial Prestressing
crete Structures Using Creep-Transformed Section Proper- for Continuous Post-Tensioned Structures in North Amer-
ties,” PCI Journal, V. 27, No. 2, Jan.-Feb., pp. 98-118. ica,” PCI Journal, V. 30, No. 1, Jan.-Feb., pp. 154-172.
Elbadry, M. M., and Ghali, A., 1989, “Serviceability Freyssinet, E., 1933, “New Ideas and Methods,” Science
Design of Continuous Prestressed Concrete Structures,” et Industrie, Jan. (See also in Trauvaux, Apr.-May 1966, pp.
PCI Journal, V. 34, No. 1, Jan-Feb, pp. 54-91. 607-622.)
Elliott, A. L., 1972, “Hindsight and Foresight on the Per- Gergely, P., and Lutz, L. A., 1968, “Maximum Crack
formance of Prestressed Concrete Bridges in the San Width in Reinforced Concrete Flexural Members,” Causes,
Fernando Earthquake,” PCI Journal, V. 17, No. 2, Mar.- Mechanisms and Control of Cracking in Concrete, SP-20,
Apr., pp. 8-16. R. E. Philleo, ed., American Concrete Institute, Farmington
El-Sheikh, M. T.; Sause, R.; Pessiki, S.; and Lu, L-W., Hills, Mich., pp. 87-117.
1999, “Seismic Behavior and Design of Unbonded Post- Ghali, A., 1989, “Stress and Strain Analysis in Prestressed
Tensioned Precast Concrete Frames,” PCI Journal, V. 44, Concrete,” PCI Journal, V. 34, No. 6, Nov.-Dec., pp. 80-95.
No. 3, May-June, pp. 54-69. Ghali, A., and Favre, R., 1986, Concrete Structures:
Emperger, F. V., 1939, “Reinforced Concrete with Addi- Stresses and Deformations, Chapman and Hall, London and
tions of High-Strength Pretensioned Steel,” Forschingsar- New York, 352 pp.
beiten auf dem Gebiete des Eisenbetons, H. 47, Wilhelm Ghali, A., and Tadros, M.K., 1985, “Partially Prestressed
Ernest und Sohn, Berlin. Concrete Structures,” Journal of Structural Engineering,
European Committee for Standardization (CEN), 1991, ASCE, V. 111, No. 8, Aug, pp. 1846-1865.
“Eurocode 2 (EC2): Design of Concrete Structures, Part 1: Giannini, R.; Menegotto, M.; and Nuti C., 1986 “Influ-
General Rules and Rules for Buildings,” Brussels, Belgium. ence of Prestressing on Seismic Response of Structures: A
Federal Emergency Management Agency, 1998, NEHRP Numerical Study,” Partial Prestressing, from Theory to
Recommended Provisions for Seismic Regulations for New Practice, NATO ASI Series, V. 2, Martinus Nijhoff Publish-
Buildings and Other Structures, Part 1 — Provisions ers, Dordrecht, The Netherlands, pp. 255-274.
(FEMA 302), Part 2—Commentary (FEMA 303), Building Guyon, Y., 1965, “Energy Absorption of Prestressed Con-
Seismic Safety Council, Feb. crete,” Proceedings, 3WCEE, New Zealand, V. III, Jan., pp.
FIP Commission on Model Code, 1984, “Recommenda- IV.216-IV.223.
tions for Practical Design of Reinforced and Prestressed Gylltoft, K., 1978, “Fatigue Tests of Concrete Sleepers,”
Concrete Structures Based on the CEB-FIP Model Code Report No. 1978:13, Division of Structural Engineering, Uni-
(MC78),” Federation Internationale de la Precontrainte versity of Lulea, Sweden, 26 pp.
(FIP), London, England. Hafner, M., and French, C. W., 1986, “Moment Resistant
FIP Commission on Prestressing Steel and Systems, 1976 Connections between Precast Elements Subjected to Cyclic
PARTIALLY PRESTRESSED CONCRETE 423.5R-33

Lateral Loads,” Structural Research Report No. 87-09, Department of Civil Engineering, West Virginia University.
Department of Civil and Mineral Engineering, University of Kent, D. C., and Park, R., 1966, “Flexural Members with
Minnesota, Minneapolis, Minn. Confined Concrete,” Journal of the Structural Division,
Harajli, M. H., 1985, “Deformation and Cracking of Par- ASCE, V. 97, No. ST5, Oct., pp. 121-146 (see also PCA
tially Prestressed Beams Under Static and Cyclic Fatigue Bulletin D108).
Loading,” PhD thesis, Department of Civil Engineering, Kunze, W. E.; Sbarounis, J. A.; and Amrhein, J. E., 1965,
University of Michigan. “Behavior of Prestressed Concrete Structures During the
Harajli, M. H., and Naaman, A. E., 1984, “Deformation Alaskan Earthquake,” PCI Journal, V. 10, No. 2, Apr., pp.
and Cracking of Partially Prestressed Concrete Beams under 80-91.
Static and Cyclic Fatigue Loading,” Report No. UMEE Lambotte, H., and Van Nieuwenburg, D., 1986, “Long-
84R1, Department of Civil Engineering, University of Term Behavior of Partially Prestressed Beams,” Partial Pre-
Michigan. stressing, From Theory to Practice, NATO ASI Series, V. 2,
Harajli, M. H., and Naaman, A. E., 1985a, “Static and Martinus Nijhoff Publishers, Dordrecht, The Netherlands.
Fatigue Tests on Partially Prestressed Beams,” Journal of Lenschow, R. J., 1986, “Partial Prestressing and Repeated
Structural Engineering, ASCE, V. 111, No. 7, July, pp. Loads Related to Random Dynamic Loading,” Partial Pre-
1602-1618. stressing, from Theory to Practice, NATO ASI Series, V. 2,
Harajli, M. H., and Naaman, A. E., 1985b, “Evaluation of Martinus Nijhoff Publishers, Dordrecht, The Netherlands,
Ultimate Steel Stress in Partially Prestressed Flexural Mem- pp. 221-229.
bers,” PCI Journal, V. 30, No. 5, Sept.-Oct. Lin, T. Y., 1963, “Load-Balancing Method for Design and
Harajli, M. H., and Naaman, A. E., 1985c “Evaluation of Analysis of Prestressed Concrete Structures,” ACI JOUR-
the Inelastic Behavior of Partially Prestressed Concrete NAL, Proceedings V. 60, No. 6, June, pp. 719-742.
Beams,” Report No. UMCE 85-2, Dept. of Civil Engineer- Lin, T. Y., and Burns, N. H., 1981, Design of Prestressed
ing, University of Michigan, 191 pp. Concrete Structures, 3rd Edition, John Wiley & Son, N.Y.
Harajli, M. H., and Naaman, A. E., 1989, “Cracking in Loov, R. E., 1987, Reader Comments on “Analysis of
Partially Prestressed Beams under Static and Cyclic Fatigue Ductility in Partially Prestressed Flexural Members,” by A.
Loading,” Cracking in Prestressed Concrete Structures, SP- E. Naaman; M. H. Harajli; and J. K. Wight, PCI Journal, V.
113, G. T. Halvorsen, and N. H. Burns, eds., American Con- 32, No. 1, Jan.-Feb., pp. 141-145.
crete Institute, Farmington Hills, Mich., pp. 29-54. Lovegrove, J. M., and El Din, S., 1982, “Deflection and
Hawkins, N. M., 1977, “Seismic Resistance of Prestressed Cracking of Reinforced Concrete under Repeated Loading
and Precast Concrete Structures —Part I,” PCI Journal, V. 22, and Fatigue,” Fatigue of Concrete Structures, SP-75, S. P.
No. 6, Nov.-Dec., pp. 80-110. Shah, ed., American Concrete Institute, Farmington Hills,
Hilsdorf, H. K., and Kesler, C. E., 1966, “Fatigue Strength Mich., pp. 133-152.
of Concrete Under Varying Flexural Stresses,” ACI JOUR- MacGregor, J. G., 1974, “Ductility of Structural Ele-
NAL, Proceedings V. 63, No. 10, Oct., pp. 1059-1075. ments,” Handbook of Concrete Engineering, First Edition,
Hsu, T.C., 1968, “Torsion of Structural Concrete Behavior M. Fintel, ed., Prentice Hall, New York.
of Reinforced Concrete Rectangular Members,” Torsion of MacGregor, J. G., and Ghoneim, M. G., 1995, “Design for
Structural Concrete, SP-18, G. P. Fisher, ed., American Torsion,” ACI Structural Journal, V. 92, No. 2, Mar.-Apr.,
Concrete Institute, Farmington Hills, Mich., pp. 261-306. pp. 211-218.
Inomata, S., 1986, “Partially Prestressed Concrete under Martin, L. D., 1977, “A Rational Method for Estimating
Cyclic Actions—Analytical Moment-Curvature Model,” Camber and Deflections,” PCI Journal, V. 22, No. 1, Jan.-
Partial Prestressing, From Theory to Practice, NATO ASI Feb. pp. 100-108.
Series, V. 2, Martinus Nijhoff Publishers, Dordrecht, The Martino, N. E., and Nilson, A. H., 1979, “Crack Widths in
Netherlands, pp. 193-204. Partially Prestressed Concrete Beams,” Report No. 79-3,
Institution of Structural Engineers., 1951, First Report on Department of Structural Engineering, Cornell University,
Prestressed Concrete, London, 285 pp. Ithaca, New York, 86 pp.
Iyengar, K. T. S. R.; Desayi, P.; and Reddy, K. N., 1970, Mattock, A. H., 1964, “Rotational Capacity of Hinging
“Stress-Strain Characteristics of Concrete Confined in Steel Region in Reinforced Concrete Beams,” Proceedings, Inter-
Binder,” Magazine of Concrete Research, V. 22, Sept., pp. national Symposium on Flexural Mechanics of Reinforced
173-184. Concrete, SP-12, American Concrete Institute, Farmington
Iwan, W. D., 1973, “A Model for the Dynamic Analysis of Hills, Mich., pp. 143-181. (See also PCA Bulletin D101.)
Deterioration Structures,” (Paper 222) Presented at the Fifth Mattock, A. H., 1967, Discussion on “Rotational Capacity
World Conference on Earthquake Engineering, held at of Concrete Beams,” Journal of Structural Division, ASCE,
Rome, Italy. V. 93, No. ST-2, Apr., pp. 519-522.
Jayashankar, V., and French, C. W., 1987, “An Interior McCall, J. T., 1958, “Probability of Fatigue Failure of
Moment Resistant Connection between Precast Elements Plain Concrete,” ACI JOURNAL, Proceedings V. 55, No. 2,
Subjected to Cyclic Lateral Loads,” Structural Research Aug., pp. 221-231.
Report No. 87-10, Department of Civil and Mineral Engi- Meier, S. W., and Gergely, P., 1981, “Flexural Crack Width
neering, University of Minnesota, Minneapolis, Minn. in Prestressed Concrete Beams,” Journal of the Structural
Jenny, D. P., 1985, Current Status of Partial Prestressing Division, ASCE, V. 107, No. ST2, Feb., pp. 429-433.
for Pretensioned Concrete Products in North America,” PCI Ministry of Transportation and Communications, 1983,
Journal, V. 30, No. 1, Jan.-Feb., pp. 142-152. Ontario Highway Bridge Design Code, Downsview,
Jittawait, J., and Tadros, M. K., 1979, “Deflection of Par- Ontario, Canada.
tially Prestressed Concrete Members,” Research Report, Moustafa, S. E., 1977 “Design of Partially Prestressed
423.5R-34 ACI COMMITTEE REPORT

Flexural Members,” PCI Journal, V. 22, No. 3, May-June. Naaman, A. E., and Hamza, A. M., 1993, “Prestress
Moustafa, S. E., 1986, “Nonlinear Analysis of Reinforced Losses in Partially Prestressed High Strength Concrete
and Prestressed Concrete Members,” PCI Journal, V. 31, Beams,” PCI Journal, V. 38, No. 3, May-June, pp. 98-114.
No. 5, Sept-Oct. Naaman, A. E.; Harajli, M. H.; and Wight, J. K., 1986,
Muguruma, H., 1986 “Seismic Problems of PPC Building “Analysis of Ductility in Partially Prestressed Concrete
Structures with Special Reference to Basic Research in Flexural Members,” PCI Journal, V. 31, No. 3, May-June,
Japan,” Partial Prestressing, From Theory to Practice, pp. 64-87.
NATO ASI Series, V. 2, Martinus Nijhoff Publishers, Dor- Naaman, A. E., and Siriaksorn, A., 1979, “Serviceability-
drecht, The Netherlands, pp. 231-254. Based Design of Partially Prestressed Beams: Part 1—Ana-
Muguruma, H., and Tominaga, H., 1972, “Deformation lytical Formulation,” PCI Journal, V. 24, No. 2, Mar.-Apr.,
Characteristics of Prestressed Lightweight Concrete Beam- pp. 64-89.
Column Assemblies under Seismic Force,” Proceedings, Naaman, A. E., and Siriaksorn, A., 1982 “Reliability of
18th National Symposium on Bridge and Structural Engi- Partially Prestressed Beams at Serviceability Limit States,”
neering, Japan, pp. 123-136. PCI Journal, V. 27, No. 6, Nov.-Dec., pp. 66-85.
Muguruma, H.; Watanabe, F.; and Fujii, M., 1983, “Shear Nadai, A., 1950, Theory of Flow and Fracture of Solids,
Behavior of Prestressed Reinforced Concrete Beams,” McGraw-Hill, New York.
Transactions, Japan Concrete Institute, V. 5, pp. 225-230. Nakano, K., 1965, “Experiments on Behavior of Pre-
Muguruma, H.; Watanabe, F.; and Fukai, S., 1980, stressed Concrete Four-Story Model Structure on Lateral
“Behavioor of Class 3 Partially Prestressed Concrete Force,” Proceedings, 3WCEE, New Zealand, V. III, Jan.,
Beam under Reversed Cyclic High Over-Load,” Proceed- pp. IV.572-IV.590.
ings of FIP Symposium on Partial Prestressing (Buchar- Nakano. K., and Okamota, S., 1978, “Test Results on
est), Federation Internationale de la Precontrainte (FIP), Beam-Column Assemblies,” Proceedings, 8th FIP Congress,
London, Sept., pp. 118-127. London, May, pp. 58-69.
Muguruma, H.; Watanabe, F.; and Fukai, S., 1983, “On the Nawy, E. G., and Chiang, J. Y., 1980, “Serviceability
Class 3 Prestressed Concrete Member with High Strength Behavior of Post-Tensioned Beams,” PCI Journal, V. 25,
Concrete,” (in Japanese), Proceedings, Annual Meeting of No. 1, Jan.-Feb., pp. 74-95.
Japan Concrete Institute, Tokyo, June, pp. 449-452. Nawy, E. G., and Huang, P. T., 1977 “Crack and Deflec-
Muguruma, H.; Watanabe, F.; and Nasu, T., 1982a, “On tion Control of Pretensioned Prestressed Beams,” PCI Jour-
the Fundamental Behaviors of a PPC Beam under Mono- nal, Vol. 22, No. 3, May-June, pp. 30-47.
tonic and Cyclic High Overload,” Journal, Japanese Pre- Nawy, E. G., and Potyondy, J. G., 1971, “Flexural Crack-
stressed Concrete Engineering Association, V. 24, Sept., pp. ing Behavior of Pretensioned Prestressed Concrete I- and T-
101-108. Beams,” ACI JOURNAL, V. 68, No. 5, May, pp. 355-360.
Muguruma, H.; Watanabe, F.; and Tanaka, H., 1982b, Nilson, A. H., 1976, “Flexural Stresses After Cracking in
“Improving the Flexural Ductility of Prestressed Concrete Partially Prestressed Beams,” PCI Journal, V. 21, No. 4,
by Using High Yield Strength Lateral Hoop Reinforce- July-August, pp. 72-81.
ment,” Journal, Japanese Prestressed Concrete Engineering Nilson, A. H., 1987, Design of Prestressed Concrete,
Association , V. 24, Sept., pp. 109-123. John Wiley & Son, N.Y.
Naaman, A. E., 1977, “Ultimate Analysis of Prestressed Nordby, G. M., 1958, “Fatigue of Concrete—A Review of
and Partially Prestressed Sections by Strain Compatibility,” Research,” ACI JOURNAL, Proceedings V. 55, No. 2, Aug.,
PCI Journal, V. 22, No. 1, Jan.-Feb., pp. 32-51. pp. 191-219.
Naaman, A. E., 1982a, Prestressed Concrete Analysis and Oertle, J., 1988, “Reibermudung einbetonierter Spannka-
Design, McGraw Hill, N.Y, 670 pp. bel (Fretting Fatigue of Bonded Post-Tensioning Tendons),”
Naaman, A. E., 1982b, “Fatigue in Partially Prestressed Report No. 166, thesis, Institute for Structural Engineering,
Beams”, Fatigue of Concrete Structures, SP-75, S. P. Shah, ETH, Zurich, Switzerland, 213 pp.
ed., American Concrete Institute, Farmington Hills, Mich., Okamoto, S., 1980, “Experimental Study on the Ductility
pp. 25-46. of PPC Beams,” Proceedings, FIP Symposium on Partial
Naaman, A. E., 1983a, “Time-Dependent Deflection of Prestressing (Bucharest), Federation Internationale de la
Prestressed Beams by the Pressure-Line Method,” PCI Precontrainte (FIP), London, Sept., pp. 128-149.
Journal, V. 28, No. 2, Mar.-Apr., pp. 98-119. Ozaka, T., 1968, “Report on Behavior of Concrete Struc-
Naaman, A. E., 1983b, “An Approximate Nonlinear Design tures During the Tokachioki Earthquake,” Journal, Japan
Procedure for Partially Prestressed Concrete Beams,” Com- Prestressed Concrete Engineering Association, V. 10, No. 5,
puters and Structures, V. 17, No. 2, pp. 287-293. Dec.
Naaman, A. E., 1985, “Partially Prestressed Concrete: Palmieri, L.; Saqan, E.; French, C.; and Kreger, M., 1996,
Review and Recommendations,” PCI Journal, V. 30, No. 6, “Ductile Connections for Precast Concrete Frame Systems,”
Nov-Dec, pp. 30-71. Mete A. Sozen Symposium: Tribute from his Students, SP-
Naaman, A. E., 1989, “Fatigue of Reinforcement in Par- 162, J. K. Wight, and M. E. Kreger, eds., American Concrete
tially Prestressed Beams,” Proceedings, ASCE Structures Institute, Farmington Hills, Mich., pp. 313-356.
Congress, San Francisco, Volume on Structural Materials, J. Park, R., and Falconer, T. J., 1983, “Ductility of Pre-
F. Orofino, ed., May, pp. 377-381. stressed Concrete Piles Subjected to Simulated Seismic
Naaman, A. E., and Founas, M., 1991, “Partially Pre- Loading,” PCI Journal, V. 28, No. 5, Sept.-Oct., pp. 111-
stressed Beams under Random Amplitude Fatigue Load- 144.
ing,” Journal of Structural Engineering, ASCE V. 117, No. Park, R., and Paulay, T., 1975, Reinforced Concrete Struc-
12, Dec., pp. 3742-3761. tures, John Wiley & Sons, New York.
PARTIALLY PRESTRESSED CONCRETE 423.5R-35

Park, R., and Thompson, K. J., 1977, “Cyclic Load Tests Model for Concrete Confinement in Tied Columns,” Jour-
on Prestressed and Partially Prestressed Beam-Column nal of the Structural Division, ASCE, V. 108, No. ST12,
Joints,” PCI Journal, V. 22, No. 5, Sept.-Oct., pp. 84-111. Dec., pp. 2703-2722.
Paulson, C. Jr.; Frank, K. H.; and Breen, J. E., 1983, “A Siriaksorn, A., and Naaman, A. E., 1979, “Serviceability-
Fatigue Study of Prestressing Strand,” Research Report No. Based Design of Partially Prestressed Beams, Part 2: Com-
300-1, Center for Transportation Research, Bureau of Engi- puterized Design and Evaluation of Major Parameters,” PCI
neering Research, University of Texas at Austin. Journal, V. 24, No.3, May-June, pp. 40-60.
PCI Ad Hoc Committee on Epoxy-Coated Strand, 1993, Soda, N., and Kazuo I., 1994, “PRC Continuous Girder
“Guidelines for the Use of Epoxy—Coated Strand”, PCI Bridge with Double-Tee Section—Kannonji Viaduct,” Pre-
Journal, V. 38, No. 4, July-Aug., pp. 26-32. stressed Concrete in Japan, Japanese Prestressed Concrete
Penzien, J., 1964, “Damping Characteristics of Pre- Engineering Association, Tokyo.
stressed Concrete,” ACI JOURNAL, Proceedings V. 61, No. Soretz, S., 1965, “Fatigue Behavior of High-Yield Steel
9, Sept., pp. 1125-1148. Reinforcement,” Concrete and Constructional Engineering,
Post-tensioning Institute, 1990, Post-Tensioning Manual, London, V. 60, No. 7, July, pp. 272-280.
Phoenix, Ariz. Spencer, R. A., 1969, “Stiffness and Damping of Nine
Precast/Prestressed Concrete Institute, 1989, Design for Cyclically Loaded Prestressed Members,” PCI Journal, V.
Fire Resistance of Precast Prestressed Concrete (MNL-124- 14, No. 3, June, pp. 39-52.
89), 2nd Edition, Chicago, Ill. Standards Association of Australia, 1984, Draft Unified
Precast/Prestressed Concrete Institute, 1992, PCI Design Concrete Structures Code, BD/2/84-11.
Handbook, 4th Edition, Chicago, Ill. Standards Association of New Zealand (NZA), 1980,
Priestley, M. J. N., and Tao, J. R., 1993, “The Seismic Code of Practice for General Structural Design and Design
Response of Precast-Prestressed Frames with Partially Deb- Loading for Buildings, (NZA 4203).
onded Tendons,” PCI Journal, V. 38, No. 1, Jan.-Feb., pp. Stelson, T. E., and Cernica, J. N., 1958, “Fatigue Proper-
58-69. ties of Concrete Beams,” ACI JOURNAL, Proceedings V. 55,
Ramirez, J.A., 1994, “Strut-Tie Design of Pretensioned No. 2, Aug., pp. 255-259.
Members,” ACI Structural Journal, V. 91, No. 5, Sept.-Oct., Sutherland, N. M., 1965, “Prestressed Concrete Earth-
pp. 572-578. quake-Resistant Structures: Development, Performance, and
Ramirez, J.A., and Breen, J.E., 1991, “Evaluation of a Current Research,” Proceedings, 3WCEE, New Zealand, V.
Modified Truss-Model Approach for Beams in Shear,” ACI III, Jan., pp. IV-463-507.
Structural Journal, V. 88, No. 5, Sept.-Oct., pp. 562-571. Swiss Society of Engineers and Architects, 1968, SIA
Rehm, G., 1960, “Contributions to the Problem of the Standard 162, Switzerland.
Fatigue Strength of Steel Bars for Concrete Reinforce- Tadros, M. K., 1982, “Expedient Service Load Analysis
ment,” Preliminary Publication, 6th Congress of the IABSE of Cracked Prestressed Concrete Sections,” PCI Journal, V.
(Stockholm), International Association for Bridge and 27, No. 6, Nov-Dec., pp. 86-111 (see also discussions and
Structural Engineering, Zurich, Switzerland, pp. 35-46. closure by H. Bachmann; E. W. Bennett; D. E. Branson; T.
Roller, J. J.; Russell, H. G.; Bruce, R. N.; and Marin, B. Brondum-Nielsen; A.S.G. Bruggeling; S. E. Moustafa; A.
T., 1995, “Long-Term Performance of Prestressed, Preten- H. Nilson; A. S. Prasada Rao; S. Natarajan; G. S.
sioned High-Strength Concrete Bridge Girders,” PCI Jour- Ramaswaym; and A. F. Shaikh; PCI Journal, Nov-Dec.
nal, V. 40, No. 6, Nov-Dec, pp. 48-59. 1983, pp. 137-158).
Rostam, S., and Pedersen, E. S., 1980, “Partially Pre- Tadros, M. K.; Ghali, A.; and Dilger, W. H., 1975, “Time-
stressed Concrete Bridges, Danish Experience,” Proceed- Dependent Prestress Loss and Deflection in Prestressed
ings, Symposium on Partial Prestressing, Bucharest, Concrete Members,” PCI Journal, V. 20, No. 3, May-June,
ICCPDC-INCERC, V. 1, pp. 361-376. (See also reprint: pp. 86-98.
Report R129, Lyngby, Technical University of Denmark, Tadros, M. K.; Ghali, A.; and Dilger, W. H., 1977, “Effect
1980.) of Non-Prestressed Steel on Prestress Loss and Deflection,”
Schroeder, A. B., 1977, “The Influence of Bond on the PCI Journal, V. 22, No. 2, Mar.-Apr., pp. 50-63.
Transfer of Cyclically Reversing Shear Across a Concrete to Tadros, M. K.; Ghali, A.; and Meyer, A. W., 1985, “Pre-
Concrete Interface,” MSCE thesis, University of Washing- stress Loss and Deflection of Precast Concrete Members,”
ton, Seattle. PCI Journal, V. 30, No. 1, Jan-Feb, pp. 114-141.
Scott, B. D.; Park, R.; and Priestley, M. N., 1982, “Stress- Taerwe, L., 1990, “Partially Prestressed Concrete Bridges
Strain Behavior of Concrete Confined by Overlapping in Belgium,” Annales des Travaux Publics De Belgique No.
Hoops at Low and High Strain Rates,” ACI JOURNAL, Pro- 5, Brussels.
ceedings V. 79, No. 1, Jan.-Feb., pp. 13-27. Tarzikhan, C., and French, C. W., 1987, “Welded and
Shahawi, M., and Batchelor, B., 1986, “Fatigue of Par- Composite Connections between Precast Elements Sub-
tially Prestressed Concrete,” Journal of Structural Engi- jected to Cyclic Lateral Loads,” Structural Research Report
neering, ASCE, V. 112, No.3, Mar., pp. 524-537. No. 87-08, Department of Civil and Mineral Engineering,
Shaikh, A. F., and Branson, D. E., 1970, “Non-Tensioned University of Minnesota, Minneapolis, Minn.
Steel in Prestressed Concrete Beams,” PCI Journal, V. 15, Tertea, I., and Onet, T., 1983, “Ductility of Partially Pre-
No. 1, Feb., pp. 14-36. stressed Concrete,” International Symposium on Nonlinear-
Sheikh, S. A., 1982, “A Comparative Study of Confine- ity and Continuity in Prestressed Concrete, Preliminary
ment Models,” ACI JOURNAL, Proceedings V. 79, No. 4, Publication, V.1, University of Waterloo, Waterloo, Ontario,
July-Aug., pp. 296-306. July, pp. 235-250.
Sheikh, S. A., and Uzumeri, S. M., 1982, “Analytical The Concrete Society, 1983, Partial Prestressing, Techni-
423.5R-36 ACI COMMITTEE REPORT

cal Report No. 23, London. Acp = area enclosed by outside perimeter of concrete
Thompson, K. J., 1975, “Ductility of Concrete Frames cross section
Under Seismic Loading,” Ph.D. thesis, University of Can- Aps = area of prestressed reinforcement in tension
terbury, Christchurch, New Zealand. zone
Thompson, K. J., and Park, R., 1980a, “Ductility of Pre- As = area of nonprestressed tension reinforcement
stressed and Partially Prestressed Concrete Beam Sections,” A′s = area of nonprestressed compression reinforcement
PCI Journal, V. 25, No. 2, Mar.-Apr., pp. 46-70. At = effective area of concrete tension zone sur-
Thompson, K. J., and Park, R., 1980b, “Seismic Response rounding all reinforcementhaving the same
of Partially Prestressed Concrete,” Journal of the Structural centroid as reinforcement
Division, ASCE, Aug., pp. 1755-1775. b = width of compression face of member
Thurlimann, B., 1971, “A Case for Partial Prestressing,” B = distance ratio at first loading cycle
Structural Concrete Symposium Proceedings, University of BN = distance ratio at the Nth loading cycle
Toronto, May, pp. 253-301. c = depth to neutral axis from extreme compression
Thurlimann, B., 1979, “Torsional Strength of Reinforced fiber
and Prestressed Concrete Beams—CEB Approach,” Con- cN = depth of neutral axis after Nth cycle of loading
crete Design: U.S. and European Practices, SP-59, Ameri- c2 = constant
can Concrete Institute, Farmington Hills, Mich., 346 pp. Cc(t–tA) = creep coefficient of concrete at time t when
Transportation Research Board, 1993, Final Draft LRFD loaded at time tA
Specification for Highway Bridge Design and Commentary, d = distance from extreme compression fiber to cen-
NCHRP 12-33, Washington, D.C. troid of nonprestressedtension reinforcement
Vecchio, J., and Collins, M.P., 1986 “Modified Compres- dc = cover to center of reinforcing steel
sion Field Theory for Reinforced Concrete Elements Sub- d′ c = concrete clear cover
jected to Shear,” ACI Journal, V. 83, No. 2, Mar.-Apr., pp. dctf = depth to centroid of tensile force in steel
219-231. de = depth to centroid of nonprestressed tension
Venuti, W. J., 1965, “A Statistical Approach to the Analy- reinforcement
sis of Fatigue Failure of Pre-stressed Concrete Beams,” ACI dmax,N = maximum deflection after Nth cycle
JOURNAL, Proceedings V. 62, Nov., pp. 1375-1394. dp = distance from extreme compression fiber to cen-
Wang, F.; Shah, S. P.; and Naaman, A. E., 1978, “High troid of prestressed reinforcement
Strength Concrete in Ultimate Strength Design,” Journal of er,n = average strain in concrete between the cracks
the Structural Division, ASCE,V. 104, No. ST11, Nov., pp. after N cycles of loading
1761-1773. es,n = steel strain at crack calculated using cracked
Wang, P. T., 1977, “Complete Stress-Strain Curve of Con- section elastic analysis afterN cycles of loading
crete and its Effect on Ductility of Reinforced Concrete Ec(t) = instantaneous elastic modulus of concrete at
Members,” PhD thesis, University of Illinois at Chicago, time t
257 pp. Ece(t) = equivalent elastic modulus of concrete at time t
Warner, R. F., and Hulsbos, C. L., 1966a, “Fatigue Proper- Eci = elastic modulus of concrete at time of loading or
ties of Prestressing Strand,” PCI Journal, V. 11, No. 1, Feb., transfer
pp. 32-52. Ei(t) = instantaneous elastic modulus of concrete at
Warner, R. F., and Hulsbos, C. L., 1966b, “Probable time t
Fatigue Life of Prestressed Concrete Beams,” PCI Journal, EN = cyclic modulus of elasticity after N cycles
V. 11, No. 2, Apr., pp. 16-39. Es = modulus of elasticity of steel
Watanabe, F.; Muguruma, H.; Tanaka, H.; and Katsuda, fc′ = concrete compressive strength
S., 1980, “Improving the Flexural Ductility of Concrete fcr = maximum recommended stress range for concrete
Beams by Using High Yield Strength Lateral Hoop Rein- ff = safe stress range
forcement,” Proceedings, FIP Symposium on Partial Pre- fmin = minimum applied stress (in concrete or reinforcing
stressing (Bucharest), Federation Internationale de la steel, as applicable)
Precontrainte (FIP), London, Sept. pp. 398-406. fpc = average precompression in concrete due to pre-
Watcharaumnuay, S., 1984, “Deflection of Partially Pre- stress at centroid of cross-section
stressed Concrete Beams Under Sustained and Cyclic Load- fpe = effective stress in prestressing steel
ing,” PhD thesis, University of Illinois at Chicago. fps = stress in prestressed reinforcement at nominal
Watcharaumnuay, S., and Naaman, A. E., 1985, “Time bending strength
and Cyclic Dependent Deflections in Partially Prestressed fpu = ultimate strength of the prestressing steel
Concrete Beams,” Report No., Department of Civil Engi- fs = tensile stress in reinforcing steel, or change in
neering, University of Michigan. steel stress from decompression
Wight, J. K., and Sozen, M. A., 1975, “Strength Decay of fso = steel stress at primary crack at first cycle
RC Columns Under Shear Reversals,” Journal of the Struc- fso,N = steel stress at primary crack at cycle N
tural Division, ASCE, V. 101, No. ST5, May, pp. 1053- fsr = maximum recommended stress range in reinforcing
1065. steel (prestressedor nonprestressed, as applica-
ble)
APPENDIX—NOTATIONS fy = yield strength of nonprestressed reinforcement
a = depth of equivalent rectangular compressive F = prestressing force
stress block F(t) = prestressing force at time t
acs = crack spacing h = depth of member
PARTIALLY PRESTRESSED CONCRETE 423.5R-37

h1 = distance from neutral axis to tension face Qser = applied service load
h2 = distance from neutral axis to centroid of r/h = ratio of base radius to height of rolled-on trans-
reinforcement verse deformation (value of 0.3 may be used in
Icr = moment of inertia of cracked section absence of specific data)
Ieff(t) = effective moment of inertia of cracked section at So = slip of reinforcement at first cycle
time t So,N = slip of reinforcement at Nth cycle
IeN = effective moment of inertia after N cycles t = time or age of concrete
Ig = gross moment of inertia tA = age of concrete at time of loading
KD, KF = constants depending on type of loading and Vc = concrete contribution of shear strength
steel profile W = crack width at the first loading cycle
Lp = equivalent plastic hinge length WN = crack width at the Nth loading cycle
M = sustained external moment at midspan Wmax = maximum crack width
Ma = applied moment Wmax,N = maximum crack width after N cycles
MD = dead load moment Z = slope of descending portion of concrete com-
Mdec = decompression moment (moment which pro- pression curve
duces zero concrete stressat extreme fiber of a β1 = parameter relating neutral axis depth to depth of
section, nearest to centroid of the prestressing equivalent rectangular compression block
force, when added to action of effective pre- ∆fps = change in stress in prestressing steel
stress alone) ∆(t) = deflection at time t
Mcr = cracking moment λ = degree of prestress
ML = live-load moment ε = strain (in general)
Mn = total nominal moment capacity ρ = reinforcement ratio of tensile nonprestressed
Mnp = moment capacity provided by prestressed reinforcement
reinforcement ρ′ = reinforcement ratio of compressive nonpre-
N = number of load cycles stressed reinforcement
Nf = number of cycles to failure ρp = reinforcement ratio of tensile prestressed
P = sum of perimeters of all tension reinforcement reinforcement
Pcp = outside perimeter of concrete cross section Σo = sum of perimeters of all tension reinforcement
PPR = partial prestressing ratio ω = reinforcing index
Q = applied load ωp = reinforcing index for prestressing steel

Potrebbero piacerti anche