Optimization of Precast Prestressed Concrete … of Precast Prestressed Concrete Bridge Girder...

19
Optimization of Precast Prestressed Concrete Bridge Girder Systems Z. Lounis, Ph.D. Research Assistant Department of Civil Enginee ri ng Un iversity of Waterloo Waterloo, Ontario, Canada M. Z. Cohn, P. Eng. Professor Department of Civil Engineering Un iversity of Waterloo Waterloo, Ontario, Canada 60 A practical approach to the optimal design of precast , prestressed concrete highway bridge girder systems is presented. The approach aims at standardizing the optimal design of bridge systems, as opposed to standardizing girder sections. Structural system optimization is shown to be more relevant than conventional girder optimization for an arbitrarily chosen structural system. Bridge system optimization is defined as the optimization of both longitudinal and transverse bridge configurations (number of spans, number of girders, girder type, reinforcements and tendon layout). As a result, the preliminary design process is much simplified by using some developed design charts from which selection of the optimum bridge system, number and type of girders, and amounts of prestressed and non-prestressed reinforcements are easily obtained for a given bridge length, width and loading type. A review of American bridge construction practice'· 2 3 enables the identifica- tion of some broad trends in the selection of bridge superstructure systems. Although this selection reflects some regional and historical variations, it appears to be primarily affected by the span length, as shown in Fig. 1. The figure suggests that reinforced and prestressed concrete solid or voided slab and light girders are suitable for short spans not exceeding 65 ft (20 m). However , for medium spans up to 130 ft (40 m), prestressed concrete stringers or box girders are preferred. For spans longer than 130 ft ( 40 m), prestressed concrete box gird- ers or other forms of construction become desirable. Fig. 1 shows that several alternative solutions exist for any given span range. For example, for spans of 65 to 100ft (20 to 30m), no less than eight candidate solutions are available. How does a designer decide on the best system selection for a given project? Bridge systems may be selected either by relying on existing tradition and accumulated experience, or by minimizing the costs under the spe- cific conditions of the project. In the first case, the solutions may not be the most PCI JOURNAL

Transcript of Optimization of Precast Prestressed Concrete … of Precast Prestressed Concrete Bridge Girder...

Page 1: Optimization of Precast Prestressed Concrete … of Precast Prestressed Concrete Bridge Girder Systems Z. Lounis, Ph.D. Research Assistant Department of Civil Engineering University

Optimization of Precast Prestressed Concrete Bridge Girder Systems

Z. Lounis, Ph.D. Research Assistant Department of Civil Engineering University of Waterloo Waterloo, Ontario, Canada

M. Z. Cohn, P. Eng. Professor Department of Civil Engineering University of Waterloo Waterloo, Ontario, Canada

60

A practical approach to the optimal design of precast, prestressed concrete highway bridge girder systems is presented. The approach aims at standardizing the optimal design of bridge systems, as opposed to standardizing girder sections. Structural system optimization is shown to be more relevant than conventional girder optimization for an arbitrarily chosen structural system. Bridge system optimization is defined as the optimization of both longitudinal and transverse bridge configurations (number of spans, number of girders, girder type, reinforcements and tendon layout). As a result, the preliminary design process is much simplified by using some developed design charts from which selection of the optimum bridge system, number and type of girders, and amounts of prestressed and non-prestressed reinforcements are easily obtained for a given bridge length, width and loading type.

Areview of American bridge construction practice'·2•

3 enables the identifica­tion of some broad trends in the selection of bridge superstructure systems. Although this selection reflects some regional and historical variations, it

appears to be primarily affected by the span length, as shown in Fig. 1. The figure suggests that reinforced and prestressed concrete solid or voided slab and light girders are suitable for short spans not exceeding 65 ft (20 m). However, for medium spans up to 130 ft (40 m), prestressed concrete stringers or box girders are preferred. For spans longer than 130 ft ( 40 m), prestressed concrete box gird­ers or other forms of construction become desirable.

Fig. 1 shows that several alternative solutions exist for any given span range. For example, for spans of 65 to 100ft (20 to 30m), no less than eight candidate solutions are available. How does a designer decide on the best system selection for a given project? Bridge systems may be selected either by relying on existing tradition and accumulated experience, or by minimizing the costs under the spe­cific conditions of the project. In the first case, the solutions may not be the most

PCI JOURNAL

Page 2: Optimization of Precast Prestressed Concrete … of Precast Prestressed Concrete Bridge Girder Systems Z. Lounis, Ph.D. Research Assistant Department of Civil Engineering University

RC SOLID SLAB

AASHTO TYPE I

AASHTO TYPE II

PC VOIDED SLAB

PCI DOUBLE T

AASHTO lYPE Ill

RC T GIRDER

AASHTO TYPE IV & PCI CHANNEL

PRECAST PCI BOX

AASHTO TYPE V

RC BOX GIRDER

AASHTO lYPE VI

RC SLAB ON PC/ STEEL GIRDER

CAST -IN ·PLACE

PC BOX GIRDER

-

SPAN LENGTH (ft)

33 tiJ 99

--I I

I I I I I

I I

10 20 30

SPAN LENGTH (m)

Fig. 1. Short and medium span bridge system selection. '·2•3

132

40

economic, while in the second case they may be very time consuming, particularly if the number of bridges to be de­signed is large.

If existing practice is to be taken as a starting point in bridge selection, the exhaustive survey< on the performance of highway bridges in the United States, published in the May-June 1992 PCI JOURNAL, can be used with great benefit. The survey indicates that since their introduction in 1950, prestressed concrete bridges have captured almost a quarter of the total bridge market, and since 1985, about 65 percent of the bridges with spans between 60 and 120 ft (18.3 and 36.6 m). The survey also indicates that, during the last 40 years, about 35,000 bridges built (or close to half of all pre stres sed concrete bridges) are of the stringer (1-girder) type (42.1 percent simple and 6 percent continu­ous stringers).

Thus, the stringer system is the most common prestressed concrete bridge type in the United States, a feature it retains throughout 25 states for spans of 60 to 80ft (18.3 to 24.4 m). It is also noted that from 1985 to 1989, prestressed concrete stringer and box girder types have each covered about one­third of the prestressed concrete bridge market.

On the other hand, if "custom-tailored" optimal designs are desired for individual bridges, designers may seek opti-

July-August 1993

165

I so

mization at any of three possible lev­els: (1 ) bridge members, (2) bridge (longitudinal and transverse) configu­rations, or (3) overall bridge system.

Level 1, member optimization of a predetermined structure, is the sim­plest and most widely reported opti­mization procedure. It involves the op­timization of girder cross sections of specified structure types such as sim­ple span prestressed concrete slabs,5

single span steel girders,6·7 single span prestressed concrete girders, 8·

9·'0·' ' sin­gle span box girders, 12 multi-span sim­ple girders, '3·14 continuous prestressed concrete box girders, '4· '

5·'6 continuous

reinforced concrete girder and frame bridges.'7

Level 2, configuration or layout op­timization, is concerned with finding the best longitudinal and transverse member arrangement within a given bridge sys tem, i.e., the number of spans, the position of intermediate supports, simply supported or continu­ous members, etc . Much less work is reported on thi s type of optimiza­tion'3·16 than on member optimization.

Level 3, system optimization, at-tempts to identify the overall features of the structural system, including ma­terial(s), structural type and configura­tion, as well as member sizing. This is the most complex design problem and only few attempt s to solve it are

known. Among these are noted the studies by McDermott, Abrams and Cohn'8•

19 on tall building systems, Gellatly and Dupree20 on surface effect vehicles , Kulka and Lin2' on medium two-span continuous prestressed concrete bridge systems, and Mafi and West7 on short simple span bridge systems.

It is apparent that the overall economic impact increases with higher optimization levels; i.e., optimizing the overall bridge system is considerably more effective in reducing total costs than optimizing its components.

Whether bridge systems are selected by following current practice or through formal optimization techniques, the de­velopment of some standards for selecting optimal bridge systems for various ranges of geometry would have consid­erable practical significance. These standard optimal sys­tems could serve both as guidelines for preliminary designs and also as yardsticks against which final bridge designs could be evaluated.

A comprehensive investigation on this topic, in progress at the University of Waterloo, aims at developing optimiza­tion procedures and solutions for short and medium span length highway bridge systems. The approach consists of first optimizing configurations of major bridge types (pre­stressed concrete stringers and box girders, steel stringers,

61

Page 3: Optimization of Precast Prestressed Concrete … of Precast Prestressed Concrete Bridge Girder Systems Z. Lounis, Ph.D. Research Assistant Department of Civil Engineering University

etc.) and then identifying optimal systems from the various pre-optimized configurations using sieve-search tech­niques.' 8· '9·20 Results on optimization of simple span bridge systems have been reported elsewhere.22

The object of this paper is to present further results on the optimization of continuous 1-girder systems and to compare them with the earlier single span solutions. More precisely, the following problem is addressed: What is the optimal pre­stressed concrete stringer system for given bridge length, width, traffic loading and standard provisions? Specifically, what are the longitudinal and transverse configurations that result in minimum superstructure cost and what are the cor­responding prestressed and non-prestressed reinforcements?

It is emphasized that "an optimal bridge superstructure" is understood as one of minimum total cost, using standard­ized girder sections and traffic loading. In this study, CPCI I-sections23 and Ontario Highway Bridge Design Code (OHBDC)24 loading, safety and design specifications, as well as Ontario costs, are used. The sensitivity of optimal solutions to the changes that occurred in the new edition of the OHBD Code25 is investigated. Alternative use of stan­dard AASHTO-PCI sections2·26 and AASHT027 specifica­tions is also considered. Worth noting is the recent interest in standardizing, across the United States, the use of pre­stressed concrete !-sections for bridge design28 which could be achieved through the formal optimization (Level 1) pro­cedures described here and elsewhere.22

This paper presents a general approach to optimization at Levels 1, 2 and 3, and readily usable results of optimal stringer configurations for precast, prestressed concrete girders. The paper is organized in three main sections: the first part details the formulation of the optimization proce­dure; the second part presents its practical implementation to bridges having design parameters within the investigation range; the third part offers some practical recommendations, extension to AASHTO specifications and design examples.

OPTIMAL BRIDGE DESIGN PROCEDURE

Design Variables

Given a bridge of total length L, width Wand some speci­fied truck and lane loads, the aim of the design is to deter­mine the minimum cost of the bridge system. The optimum bridge is defined by the best structural system type (longitu­dinal and transverse bridge configurations) and optimum de­sign of its individual members (slab, girders, diaphragms, etc.). In this paper, the selected structural system consists of a reinforced concrete slab on precast, prestressed concrete !-girders, as it represents about 40 percent of the short and medium span bridges built in the United States and Canada.4·2'

The optimum longitudinal configuration is defined by the number of spans, restraint type (simply supported or contin­uous) and span length ratios (Fig. 2). The transverse config­uration is defined by the number of girders (or girder spac­ing and slab overhang) (Fig. 3). Simply supported bridges (with one, two and three spans) and continuous bridges with two and three spans are considered, although the results are

62

conservative and may be used for four or more equal spans. Finally, the optimum designs of the girder and slab are

defined by the girder cross section dimensions (Figs. 4 and 5a), slab thickness, amounts of prestressed and non­prestressed flexural reinforcements in the girder and slab, as well as the tendon eccentricities at midspan and supports (Figs. 5a and 5b).

Objective Function

The assumed merit function is the unit superstructure cost and is defined as:

Z = [n C8 L + Ccs Vcs + C5 W, + n' Cpc + C5 W,, + C5 W,p] I W L (1)

where

n = number of girders

n' = number of positive moment connections (at piers)

C8

= cost of precast girder per length (including cost of fabrication, prestressing, delivery and erection)

Ccs = concrete cost in slab and diaphragms

C5 = non-prestressed steel unit cost

Cpc = unit cost of positive moment connection at piers including cost of steel and bending of bars (as­sumed cpc = 1.7 cs ~c· where vpc =volume of pos­itive moment reinforcement)

Vcs = volume of concrete in slab and diaphragms and W,, W.n and W,p are the weights of non-prestressed steel in slab and diaphragms, negative moment steel in slab at piers and positive moment steel in girders, respectively.

Design Constraints

In the design of precast concrete girders, all serviceability limit states (SLS) and ultimate limit states (ULS) require­ments of the Ontario Highway Bridge Design Code24

(OHBDC-1983) are to be satisfied. However, only the flex­ural constraints at SLS, transfer and ULS are considered, in addition to the side constraints that reflect specific OHBDC requirements. Other constraints (camber, deflection, vertical and interface shears, etc.) could easily be added but, in gen­eral, they affect only marginally the flexural design and are left to be checked at the final design stage. Special attention must be given to the connection design and detailing for the positive moments at piers induced by the creep and shrink­age effects.

(a) Transfer Stresses Constraints- Normal stresses in concrete O'er due to girder self-weight should be within the allowable limits24 at midspan and support critical sections:

(2)

where fu and frc are the allowable tensile and compressive stresses at transfer, respectively. For the girder with a con­crete having J; = 40 MPa (5800 psi), j,1 = -0.24 {!;; = -1.3 MPa (-190 psi) and frc= 0.6:f;; = 18 MPa (2610 psi).

(b) Service Limit State Constraints - In the design of prestressing steel for girders and non-prestressed reinforce-

PCI JOURNAL

Page 4: Optimization of Precast Prestressed Concrete … of Precast Prestressed Concrete Bridge Girder Systems Z. Lounis, Ph.D. Research Assistant Department of Civil Engineering University

1-S~ I• •I

(a)

2-S

3-S~ L/.3 ~ L/.3 ~ L/.3 ~

(c)

2-c~ ~ ~ L/2 I• L/2 •I• • I (d)

3-c ~ ~ ~ ~ L/.3 L/.3 L/.3

(e)

Fig. 2. Longitudinal bridge configurations: (a) one simple span , (b) two simple spans, (c) three simple spans, (d) two continuous spans, (e) three continuous spans.

s ·I· s ~l·s' :I 1: s' ~I· s ~I· w

Fig. 3. Transverse bridge configuration.

ment for positive moments at piers, the normal stresses in concrete (<Yes ), mild steel (as) and jacking prestress ( Oj) should be: For concrete:

fst ~ <Yes ~ f sc (3a)

where fsr = -0.48 -fJZ = -3 MPa (-435 psi) and fsc = 0.45f~ = 18 MPa (2610 psi) are the allowable24 tensile and com­pressive stresses at service, respectively. For non-prestressed steel:

(3b)

For prestressing steel at jacking:

(3c)

The SLS moment is defined as:24

July-August 1993

CPCI900 CPCI1200 CPCI1400

Fig. 4. Standard CPCI 1-girder sections.23 All dimensions in mm ; 1 mm = 0.0394 in.

(a)

C.G.Steel C. G. Concrete

~:_~:_---- -]:--- -:_~-~ ·-·- · ·--·-·

!/3

(b)

Fig. 5. Girder design variables: (a) cross section, (b) tendon layout.

:I

(4)

where M8

, Ms and Ma are the girder, slab and asphalt pave­ment self-weight moments, respectively. Also, ML and Mcs are the live load moment and creep and shrinkage moment, respectively.

(c) Ultimate Limit State Constraints - The ultimate load moment Mu should be less than or equal to the resisting moment of the section:

(5)

where Mn is the nominal resisting moment of the section and If>= 0.85 is the flexural strength reduction factor. In the new OHBD Code/' the single understrength factor (If> = 0.85) has

63

Page 5: Optimization of Precast Prestressed Concrete … of Precast Prestressed Concrete Bridge Girder Systems Z. Lounis, Ph.D. Research Assistant Department of Civil Engineering University

been replaced by strength reduction factors for concrete ( if>c = 0. 75), steel ( if>s = 0.90) and prestressing steel ( 1/>p = 0.95).

(d) Side Constraints - These reflect some specific OHBDC requirements:

s ~ 15 t (6a)

t ~ 225 mrn (8.9 in.) (6b)

S' ~Min [0.6 S, 1.8 m (5 .9 ft)] (6c)

(() ~ 0.30 (6d)

c bor ~ 100 mrn (3.9 in.) (6e)

Crop~ 80 mrn (3.1 in.) (6f)

where S, S ', t, w, cbor and C10P are the girder spacing, slab overhang, slab thickness, total tension reinforcement index of the section, and bottom and top minimum concrete cov­ers, respectively. The total reinforcement index is defined as:

(7)

Dead and Live Load Analysis

In order to properly account for the construction sequence of this type of bridge, two analytical models are used to compute the moments at different critical sections.

(a) Model 1: Simply Supported Beam - Initially, the precast girders are placed on the substructure and the analy­sis under the girder and slab self-weights is carried out as for simnle beams. At this stage, only the girder section is ef­fective in computing the stresses.

(b) Model 2: Continuous Beam- Subsequently, conti­nuity is achieved by adding non-prestressed reinforcement over the piers, and the analysis for the live load and pave­ment weight is carried out as for a continuous beam. Com­posite (slab and girder) action develops and must be consid­ered in evaluating the stresses . However, the live load moment obtained should be multiplied by the transverse dis­tribution factor S/Dd and augmented by the impact factor I (dynamic load effect) , conservatively estimated as 0.4 for truck load and 0.1 for uniform lane load. In OHBDC,24 Dd values (given for two-, three- and four-lane bridges) depend on the torsional and flexural parameters a and 8, respec­tively, in addition to the design lane width:

Dxy + Dyx + D1 + D2 a= o s

2(Dx Dy) .

( )

0.25

()=~ Dx u * v y

(8a)

(8b)

where Dx> Dy, Dxy, Dyx> D1 and D2 are the flexural , torsional and coupling rigidities in both transverse and longitudinal directions, respectively. The Dd factors have been derived by using the orthotropic plate analogy for simple span bridges. 29 In order to use the same factors for continuous

64

• .e,-o.e.t I•

I ¥1 i ¥ • • .t.2•0.4.t .t5•0.6.t

I 1'4 •I. •I

Fig. 6. Equivalent spans for live load moment computations at critical sections. 24

·30

continuity reinforcement

~-

(a)

/ reinforced concrete slab

x~ ;g F ®

I . '• -ln+1

(b)

Fig. 7. Analytical models for continuous precast girders: (a) statically determinate system, (b) statically indeterminate system.

bridges, an effective length £* is determined for all critical sections to compute the flexural parameter in Eq. (8b) which may be taken as shown24

·30 in Fig. 6. In the new edition of

the OHBD Code,25 the determination of the Dd factors has been simplified to only depend on the bridge type and span length.

Creep and Shrinkage Moments Analysis

(a) Creep Moments due to Prestressing and Dead Loads - Continuity of this bridge type requires that posi­tive moments developed over piers due to creep of the pre­stressed girders and the effects of live load on remote spans be considered. Positive moments due to creep under sus­tained (prestressing and dead) loads are partially counter­acted by the negative moments due to differential shrinkage between the cast-in-place deck slab and precast girders. Of the several methods of creep analysis proposed in the litera­ture , '1.3

233 the Trost-Bazant age-adjusted elastic modulus method33 is used to compute the prestress and dead load re­straint moments .

For illustration purposes, consider a continuous girder with n spans, initially assumed freely supported (Fig. 7a) , and subsequently connected for continuity at an age t1 (Fig. 7b). Using the age-adjusted elastic modulus method, the moment M~ (t) induced by creep at a support n is given by:

PCI JOURNAL

Page 6: Optimization of Precast Prestressed Concrete … of Precast Prestressed Concrete Bridge Girder Systems Z. Lounis, Ph.D. Research Assistant Department of Civil Engineering University

where Mn is the moment at support n if the structure is con­tinuous and monolithically cast in a single stage.

The restraint moments at the supports are determined by using any elastic analysis method (e.g., the three-moment equation); v(t, t0 ) is the creep coefficient (ratio of creep strain to initial elastic strain); t0 is the age at loading; t1 is the age at which continuity is achieved and X is the aging coefficient introduced because the creep moment develops gradually with time (Trost ptoposes X= 0.8). Assuming t1 = 28 days and t = oo, the following values are obtained from ACI-209:33 V (t, t0 ) = 0.96 Vu, V (tl, t0 ) = 0.42 Vu and V (t, t1) = 0.54 vu, where vu is the ultimate creep factor which may be conservatively taken as 4.5.

Since v(t1, t0 ) = 0.42vu (i.e., 42 percent of the ultimate creep will have occurred at the time the continuity is achieved), the connections should develop the moments from the remaining creep of 0.54 Vu· Substitution of parame­ters in Eq. (9) yields:

M~ (t = oo) = 0.80 Mn (10)

(b) Shrinkage Moments - Differential shrinkage be­tween the reinforced concrete deck slab and the prestressed concrete girders induces the moment Msh in the composite section:

(11)

where

A5 = area of effective slab deck £ 5 =Young's modulus of slab concrete Ys = distance from centroid of composite section to cen­

troid of slab t:sh = differential shrinkage strain24 = 1 0_.

The (shrinkage) restraint moments induced by differential shrinkage can also be obtained using the three-moment equation; however, the presence of creep reduces shrinkage restraint moments. In the absence of accurate data, the OHBDC24 recommends the use of a 60 percent creep reduc­tion factor. Therefore, the final moment Mcsn at support n due to the combined effects of creep and shrinkage is :

where Mpn' MDn and Mshn are the restraint moments at support n due to prestressing, dead load and shrinkage, respectively.

Formulation of the Optimal Bridge Design Problem

The optimization problem may be stated as follows : Minimize:

Z = [n CgL + Ccs ~s + Cs W, + n' Cpc + Cs W.n + Cs l-v,p] I W L (13a)

such that:

(13b)

(13c)

July-August 1993

Mj = 0.85 [0.8 (Mpj +MD).:... 0.4 Msh) + 0.75 MLj ~ Asj (0.6 fy) (d- kd/3) ( 13d)

U = 1, 2, ... np) Muj = 1.5 Maj + 1.4 MLj ~ 0.85 Mnj (13e)

(i = 1, 2, . . . n5 ) Mui = 1.1 Mgi + 1.2 M5; + 1.5 Mai + 1.4 Mu ~ 0.85 Mni (13f)

s ~ 15 t (13g)

t ~ 225 mm (8.86 in.) (13h)

S' ~Min [0.6 S, 1.8 m (5.91 ft)] (13i)

Q)~ 0.30 (13j)

cbor ~ 100 mm (3.94 in.) (13k)

Crop~ 80 mm (3.1 in.) (131)

Eqs . (13b) and (13c) represent the transfer and service stresses constraints, respectively. Eq. (13d) represents the SLS constraint on the steel stress for the design of positive moment reinforcement at the piers. Eqs. (13e) and (13f) are the ULS flexural strength constraints for negative moment reinforcement at the nP piers and positive moment reinforce­ment at the n5 critical span sections, respectively. Eqs. (13g) to (131) are the side constraints.

Solution of the Optimization Problem

The optimization problem, defined by Eqs. (13a) to (131), includes both discrete and continuous design variables; the constraints are nonlinear functions of the design variables. Solution of such a problem is very difficult and requires the use of mixed integer programming algorithms which are not efficient and may not converge to an optimum. To over­come this difficulty, a two-stage optimization with continu­ous variables only is performed by using nonlinear program­ming methods. This approach has been used for the optimization of simply supported bridge girders.22

The two-stage optimization operates as follows: (a) Maximum Feasible Girder Spacing - For each

simple span and continuous bridge and precast girder, the maximum feasible girder spacing is determined for span lengths varying from 10 to 30 m (33 to 98 ft) by solving the optimization problem defined by Eqs. (13a) to (131), but maximizing the girder spacing instead of minimizing the cost as objective function. Once the maximum feasible girder spacing is determined, for every given bridge width and span length, the minimum number of girders resulting in the minimum superstructure cost can be obtained.

(b) Optimum Reinforcements - After the optimum number of girders has been determined, the total superstruc­ture cost can be minimized by solving the problem defined by Eqs. (13a) to (131) with the minimization of the pre­stressed and non-prestressed reinforcements as objective function.

These two optimization problems can be solved by any nonlinear programming technique. In this paper, the

65

Page 7: Optimization of Precast Prestressed Concrete … of Precast Prestressed Concrete Bridge Girder Systems Z. Lounis, Ph.D. Research Assistant Department of Civil Engineering University

GAMS/MINOS program, based on the projected Lagrangian algorithm, has been used. The description of the algorithm is given elsewhere.34

•35

OPTIMAL DESIGN OF BRIDGE SYSTEMS­PRACTICAL APPLICATION

Investigation Outline

To achieve the stated objectives of the investigation, the above-mentioned optimization procedure and the GAMS/MINOS program have been used to produce optimal bridge designs for a large number of precast, prestressed con­crete girder bridges. Some 165 optimal solutions have been generated for the following variations of the main design data indicated by the dots (Table 1):

• Optimal prestressing (force and tendon layout) • Non-prestressed steel reinforcements for girder and slab • Minimum cost of superstructure per unit deck area • Identification of active constraints

Optimal Design Solutions for Simple and Two- and Three- Span Continuous CPCI Girders

Optimal solution features for the simple and two- and three-span continuous girders are presented in Table 2 and Figs. 8 and 9. With the completed optimal designs for the simple and two- and three-span continuous girders, the op­timal superstructure costs for discrete bridge lengths up to 100 m (330 ft), three bridge widths and three CPCI girder types in five longitudinal bridge configurations are summa­rized in Table 3.

(a) Five longitudinal bridge configu­rations: one, two and three simply sup­ported spans, and two and three con­tinuous equal spans

Table 1. Specimen data for parametric investigation for bridge widths W= 8, 12 and 16m (26, 39 and 52ft).

(b) Three girder types : CPCI 900, 1200 and 1400

(c) Three bridge widths: W = 8, 12 and 16 m (26, 39 and 52 ft) (i.e., two-, three- and four-lane bridges)

(d) Five span lengths: 10, 15, 20, 25 and 30 m (33, 49, 66, 82 and 98 ft) (whenever feasible)

For all designs, the following values are adopted as pre-assigned parameters: • Class A highway • Slab thickness: t = 225 mm (8.9 in .)

(OHBDC minimum) • Minimum OHBDC slab reinforce­

ment (i.e., p = 0.3 percent top and bottom reinforcements in both directions)

• Concrete grades: f~ = 40 MPa (5800 psi); f~; = 30 MPa (4350 psi) for girder and f~ = 30 MPa (4350 psi) for slab

• Steel grades: ipu = 1860 MPa (270 ksi) for prestressing steel and !y = 400 MPa (60 ksi) for mild steel

• Unit costs based on average Ontario costs always expressed in Canadian dollars: cg = $514, 486, 580 perm for CPCI 900, 1200 and 1400, re­spectively; Ccs = $617 per m3 and C5

= $1400 pert For simple span bridges, Eq . (1)

with cpc = WSII = 0 applies, but the additional cost of deck joints, c dj = $1410 perm, must be added.

Solutions of the 165 bridge designs include: • Optimal girder spacing and the cor­

responding minimum number of girders

66

Span Longitudinal bridge configuration and CPCI girder type

length 1-S 2-S 3-S 2-C 3-C

(m) 900 1200 1400 900 1200 1400 900 1200 1400 900 1200 1400 900 1200 1400

10 • • • • • • • • • • • • • • • 15 • • • • • • • • • • • • • • • 20 • • • • • • • • • • 25 • • • • • • • • • • 30 • • • • •

Metric (SI) conversion factor: I m = 3.28 ft.

Table 2. Maximum feasible girder spacing, optimum girder spacing and optimum number of girders for single-, two- and three-span continuous CPCI girder bridges.

Girder Girder type and span length Bridge spacing

CPCI900 CPCI 1200 CPCI 1400 width and number W(m) of girders 10m ISm 10m 15m 20m 25m 10m 15m 20m 25m 30m

Maximum All feasible 3.37 2.52 3.37 3.37 2.95 1.95 3.37 3.37 3.37 3.10 2.19

"' widths spacing (m) " OJ)

~ S(m) 2.5 2.5 2.5 2.5 2.5 1.9 2.5 2.5 2.5 2.5 1.9 ..0 ... 8m

" 3 4 '0 n 3 3 3 3 3 4 3 3 3 ·61 ~ S(m) 2.9 2.3 2.9 2.9 2.9 1.95 2.9 2.9 2.9 2.9 2.0 ~ 12 m

" n 4 5 4 4 4 6 4 4 4 4 6 Oil c:

i:i) S (m) 3.1 2.2 3. 1 3.1 2.6 1.95 3.1 3.1 3.1 3.1 2.2 16m

n 5 7 5 5 6 8 5 5 5 5 7

"' Maximum " OJ)

~ All feasible 3.37 2.79 3.37 3.37 3. 10 2.19 3.37 3.37 3.37 3.30 2.33 ..0 widths spacing (m) il 'E ·en S(m) 2.5 2.5 2.5 2.5 2.5 1.9 2.5 2.5 2.5 2.5 1.9 "' 8m ::J 0 n 3 3 3 3 3 4 3 3 3 3 4 ::J c:

'.;:l c: S(m) 2.9 2.9 2.9 2.9 2.9 2.0 2.9 2.9 2.9 2.9 2.3 0 C,) 12m c: n 4 4 4 4 '"

4 6 4 4 4 4 5

ff 2.6 3. 1 3.1 3. 1 2.0 3.1 3.1 3.1 3. 1 2.2 !"'\ S(m) 3.1

o(l 16m M n 5 6 5 5 5 8 5 5 5 5 7

Metric (S I) conversion factor: I m = 3.28 ft.

PCI JOURNAL

Page 8: Optimization of Precast Prestressed Concrete … of Precast Prestressed Concrete Bridge Girder Systems Z. Lounis, Ph.D. Research Assistant Department of Civil Engineering University

Table 3. Minimum superstructure cost per deck area ($/m2) for investigated bridge systems.

Bridge Bridge 1-S 2-S length width L (m) W(m) 900 1200 1400 900 1200 1400

8 363 352 388 10 12 341 332 363

16 331 322 351

8 363 352 388 15 12 384 332 363

16 395 322 351

8 352 388 441 433 470 20 12 332 363 420 413 446

16 352 351 409 403 434

8 413 388 426 417 454 25 12 413 363 404 397 430

16 443 351 426 387 417

8 460 415 406 443 30 12 460 437 386 419

16 424 447 376 406

8 393 429 40 12 373 405

16 393 393

8 449 424 45 12 409 400

16 449 424

8 445 421 50 12 445 397

16 476 384

8 488 60 12 488

16 451

8 75 12

16

8 90 12

16

Metric (Sl) conversion factor: 1m' = 10.76 sq ft.

Finally, the optimization results are synthesized in Fig. 10, which identifies the optimum precast girder bridge sys­tem (i .e., the best longitudinal configuration, as well as the optimum number and type of girders) for a specified bridge length and width.

(a) Optimum Girder Spacing (Minimum Number of Girders) - Optimization results show that the optimum girder spacing is the same for two- and three-span continu­ous girders. The maximum feasible and optimum girder spacing and corresponding number of girders are shown in Table 2 for the three bridge widths [8, 12 and 16m (26, 39 and 52 ft)] investigated. From Table 2, it is seen that the maximum feasible girder spacing depends on the span length, girder type and slab thickness. However, the opti­mum girder spacing depends on the bridge width in addition to the above-mentioned parameters.

It is also noted that, for a given precast girder section, bridge span and width, the optimal girder spacing is less than the maximum feasible value and always corresponds to the maximum feasible slab overhang. For a given precast girder section and bridge width, the optimum girder spac-

July-August 1993

2-C 3-S 3-C

900 1200 1400 900 1200 1400 900 1200 1400

371 365 403 349 344 378 339 333 366

370 363 401 348 342 376 353 332 364

368 361 399 467 460 498 374 369 376 346 340 374 446 440 473 353 348 383 367 330 361 435 430 461 342 337 371

358 396 441 433 470 372 366 379 338 371 463 413 446 351 346 381 328 359 473 403 434 360 335 369

387 395 432 424 461 371 364 381 377 371 454 404 437 349 344 380 372 359 465 394 425 371 333 368

416 394 426 417 453 370 363 382 416 370 447 398 429 348 343 379 416 358 458 387 417 370 332 366

464 406 443 362 383 416 386 418 342 376 428 406 406 331 363

456 432 421 385 456 407 426 374 487 395 382 362

497 457 497 422 460 433

ing (or number of girders) is a constant that minimizes su­perstructure cost up to a certain span length. Beyond this span length, an extra girder is required, at an increased su­perstructure cost. For CPCI 1200 and 1400 girders, these span lengths are .e = 20 m (65.6 ft) and .e = 25 m (82.0 ft), respectively.

(b) Optimum Prestressing Force and Tendon Layout - Optimization results show that the optimum prestressing force and total reinforcement index for the three-span con­tinuous girders are up to 2 percent higher than for those cor­responding to the two-span continuous girders. This minor difference is due to the fact that a large part of the load (girder and slab self-weights) is carried in the same way (simple beam) by both configurations.

Further, since both resist the same dead load moments and only marginally different live load moments, the variations of the optimum prestressing force with span length for vari­ous CPCI girders and bridge widths are illustrated in Fig. 8b for the three-span continuous girders only. However, this plot can be conservatively used for continuous beams with any number of equal spans. The optimum prestressing

67

Page 9: Optimization of Precast Prestressed Concrete … of Precast Prestressed Concrete Bridge Girder Systems Z. Lounis, Ph.D. Research Assistant Department of Civil Engineering University

4000

4000 Walm

~ w =am z ~ 3000 CPCI1400

J 3000 CPCI 1400

i Oil CPCI 1200 CPCI1200 .5 Oil

I c: 2000

2000 "ii !!

CPCI 000~ l e ::I

1000 § 1000 e 8 e

! 0 0

0 10 20 30 40 0 10 20 30

Span length (m) Span length (m)

4000 4000

~ W=Um ~ WaUm

J 3000 CPCI 1400 ] 3000 Oil Oil

1 c:

CPCI1208 -CPCI 1200 li 2000 i 2000

a § e

::I

! 1000 e 1000

8 0 0

0 10 20 30 40 0 10 20 30

Span length (m) .Span length (m)

4000 4000

~ W =16m

~ Wa16m

] 3000 i 3000 Oil .5 Oil CPCIUOO

I 2000 1 2000

~ 1000 e 1000 ::I

8 .5 c!

0 0 0 10 20 30 40 0 10 20 30

Span length (m) Span length (m)

(a) (b)

Fig. 8. Optimum prestressing force for CPCI girder bridges: (a) simple span bridges, (b) continuous span bridges. Note: 1 kN = 0.225 kips; 1 m = 3.28 ft.

68

40

40

40

PCI JOURNAL

Page 10: Optimization of Precast Prestressed Concrete … of Precast Prestressed Concrete Bridge Girder Systems Z. Lounis, Ph.D. Research Assistant Department of Civil Engineering University

0.07 0.07

:s .8 ~ 0.06 CPCI .s 0.06

i!! 5 ll 0.05 CPCI 0.05 e E

j 0.04 CPCI j 0.04 .5 -~ e 0.03 0.03 5 s 9

0.02 g

0.02 § § -~ 0.01 E 0.01 ·~

8 8'

CPCI

CPCI 12001--__.

0.00 0.00 0 10 20 30 40 0 10 20 30

Span length (m) Span length (m)

(a) (b)

Fig. 9. Optimum total reinforcement index for CPCI girder bridges: (a) simple span bridges, (b) continuous span bridges (all bridge widths) . Note: 1 m = 3.28 ft.

40

forces for simply supported girders shown in Fig. Sa are seen to be 10 to 40 percent higher than for the corre­sponding continuous girders.

Table 4. Optimum number and type of girders for two- and three-span continuous CPCI girder bridges and various bridge widths.

As expected, it is seen that for a given span length, the amount of pre­stressing decreases for larger girder depths. The lowest values always cor­respond to the CPCI 1400 section, which requires very limited prestress­ing for spans smaller than 15 m (49 ft). This suggests that for short spans, a reinforced concrete girder with the same depth will be able to sustain the same loads.

Bridge

width W(m)

8

9

10

11

12

13

14

15

16

CPCI900 10 15 10 3 3 3

3 4 3

3 4 3

4 4 4

4 4 4

4 5 4

4 5 4

5 6 5

5 6 5

Girder type and span length, m

CPCI1200 CPCI 1400 15 20 25 10 15 20 25 30 3 3 4 3 3 3 3 4

3 3 4 3 3 3 3 4

3 3 5 3 3 3 3 5

4 4 6 4 4 4 4 5

4 4 6 4 4 4 4 5

4 4 6 4 4 4 4 6

4 5 7 4 4 4 5 6

5 5 7 5 5 5 5 7

5 5 8 5 5 5 5 7 The optimum prestressing force al­ways corresponds to the maximum midspan tendon eccentricity (i.e., min-

Metrtc (SI) converston factor: l m = 3.28 ft.

imum concrete cover), with the tensile service stress con­straint being active. However, for the CPCI 900 section and f = 15 m (49ft), the tendon eccentricity is not maximal be­cause the transfer tensile stress constraint also becomes ac­tive and thus limits the feasible eccentricity. For the end support eccentricity, any feasible solution can be used since its effect on the optimum design is negligible. However, tak­ing the smallest feasible value may reduce the amount of positive moment reinforcement at the piers.

(c) Optimum Total Reinforcement Index- The total reinforcement index is found to be rather low (comax = 0.06) for the three investigated girder sections. This suggests that the ductility constraint [Eq. (6d)] is redundant for this type of girder. Furthermore, the identical values obtained for all bridge widths indicate that the number of lanes does not in­fluence the total reinforcement index.

Some additional non-prestressed reinforcement is re­quired in compliance with the ULS flexural strength crite­rion. This reinforcement percentage varies between 0.85 and 0 percent, with higher percentages needed for shorter spans. Variations of co with span length for the three precast girder

July-August 1993

types are shown in Fig. 9. Since the amount of prestressing is known from Fig. 8, and the total reinforcement index is known from Fig. 9, the amount of non-prestressed rein­forcement is easily determined from Eq. (7) .

(d) Optimum Unit Cost of Superstructure- Table 3 shows unit superstructure costs (i.e., cost per deck area) for bridge lengths varying from 10 to 90 m (33 to 295 ft), three bridge widths [8, 12 and 16m (26, 39 and 52ft)], three CPCI girder types and the five longitudinal bridge configurations investigated in this paper. From Table 3 it is noted that: • Configurations 2-S and 3-S are never competitive because

of the added cost of the deck joints and are reported here for the sake of completeness only.

• For the competitive configurations (1-S, 2-C and 3-C), costs are almost constant up to a certain span length be­yond which they increase sharply, due to an increase in the required number of girders. As an example, the cost of CPCI 1200 at f =25m (82ft) increases on average by 17 percent vs. the cost at f = 20 m (66 ft).

• In general, unit cost decreases for increasing number of lanes. Thus, the superstructure cost for CPCI 1400 and

69

Page 11: Optimization of Precast Prestressed Concrete … of Precast Prestressed Concrete Bridge Girder Systems Z. Lounis, Ph.D. Research Assistant Department of Civil Engineering University

-N

E

0 ~ 0 ~ () (]) "0 -+-

0 (.)

cs~~----------------------------~ W•8m 4&) (2-lane bridge)

4&l~w~.~12~m----------------------------------~

425 (3-lane bridge)

3-C . ~--~ L/3 ot 1.(3 * L/3 .r I 5-l.(X)

2-c ~I . ~L/2 ? L/2 f J

375 l·S .... '·• ' 4-l.(X) r -~ ·~l.(X) 4-l.(X) ~1l-!--4'"' : . . I

350 4-9{1J\ : - -- _J ... ~-=:. 4-l:m 325 •• •• • •• • 4-laD

4-l:lD

&QL-__ ._ __ ._ __ ._ __ ~--~--~--~--~--~~

4&l~w~.~l6~m----------------------------------~

400

425 (4-lane l:xldge) 3-C f7-1'4D ~ L/3 .'fi L/3 4!. L/3~ 1

2-<: ..J , L/2 ? 1./2 ~ \Jf6.14XJ

375 ~ ·,\· • 'I I ·--~·· 5-14D

I\ 6:1~5-l.(X) ,... 00

3eO ~~··· I . ~ .s:mr.. ____ j 325 • • • •• • 5-l:m 5-12CO

5-l:m

Total bridge length L(m)

72 ft and f > 72 ft), the CPCI 1200 and 1400 are optimal, respectively.

• For total bridge lengths L ~ 66 m (217 ft), the optimal system always consists of CPCI 1200 sections, with three, four or five girders for W = 8, 12 or 16m (26, 39 or 52ft), respectively.

• For L > 66 m (217 ft), the optimal system always consists of CPCI 1400 sections, with four to seven girders depending on W and L (Fig. 10 and Table 4).

(e) Governing Criteria (Active Constraints) - In all cases investi­gated, the service tensile stress [ <Yes = fsv Eq. (13c)] is the only active SLS constraint. Therefore, all other SLS constraints may be disregarded, except for some special cases where transfer stress constraints may also be active and need to be considered [for f = 25 and 30 m (82 and 98 ft) , <Yet = fcc, Eq. (13 b) governs for CPCI 1200 and 1400, respectively, while <Yet = fu, Eq. (13b) governs for CPCI 900 and f = 15m (49ft)].

Optimum Bridge Superstructure System

At this point, the problem to be solved is: What is the optimum system for some given bridge length and width; i.e., which of the five longitudi­nal configurations and transverse con­figurations result in the most economi­cal superstructure?

The answer is provided by Fig. 10, which is a synthesis of optimal design solutions in Tables 2 and 3. Fig. 10 il­lustrates the variations of optimal bridge costs per unit deck area (Z) with the total bridge length (L), for various longitudinal configurations and girder sections. Its study justifies the following remarks:

Fig. 10. Optimum bridge systems for W = 8, 12 and 16 m (26, 39 and 52 ft). Note: $1/m2 = 0.093 $/fF.

• Solutions 2-S and 3-S (two and three simple spans) are not plotted because they are not competitive

(high additional costs of the required deck joints). .f = 25 m (82 ft) of a fo ur-lane bridge is 10 percent cheaper than that of a two-lane bridge.

• For short span bridges [ f ~ 13 m ( 43 ft)], both CPCI 1200 and 900 may be considered equally competitive for all configurations, because CPCI 1200 is a little cheaper than CPCI 900. However, if the unit costs of the two gird­ers are equal, the CPCI 900 girder will be optimal for this span range. For 13 < .f ~22m and f >22m (43 < f ~

70

• The optimum longitudinal configuration depends only on the total bridge length and is independent of its width.

• Bridge configurations 1-S, 2-C and 3-C are optimal for short [L ~ 27 m (89 ft)], short to medium [27 < L ~ 44 m (89 < L ~ 144ft)] and medium [55 < L ~ 100m (180 < L ~328ft)] length bridges, respectively.

• For 44 < L ~55 m (144 < L ~ 180ft), the three-span con-

PCI JOURNAL

Page 12: Optimization of Precast Prestressed Concrete … of Precast Prestressed Concrete Bridge Girder Systems Z. Lounis, Ph.D. Research Assistant Department of Civil Engineering University

tinuous (Solution 3-C) bridge is not necessarily optimal although its superstructure cost is about 7.5 percent cheaper than that of the two-span continuous (Solution 2-C) bridge. Determination of the optimum configuration should be based on the minimum total bridge cost (super­structure plus substructure), as discussed in the next section.

• For 27 < L ~ 44 m (89 < L ~ 144ft), the two-span contin­uous (Solution 2-C) bridge is optimal with CPCI 900 and 1200 as girder sections. The optimum number of girders is three, four and five for W = 8, 12 and 16 m (26, 39 and 52ft), respectively, as shown in Fig. 10.

• For 55< L ~ 100m (180 < L ~328ft) , Solution 3-C is the only configuration that remains feasible. The CPCI 1200 and 1400 are the optimal girder sections for bridge lengths smaller or larger than 65 m (213ft) , respectively.

• For L > 65 m (213 ft), the three-span continuous bridge with CPCI 1400 girders should be compared to the four­span continuous bridge with CPCI 1200 or 1400 girders and other systems - such as voided slab or box girder bridges - which may be competitive for medium to long span bridges.

• For short span bridges with f ~13m (43ft) [i.e., L ~ 13m ( 43 ft) for Solution 1-S, L ~ 26 m (85 ft) for Solution 2-C and L ~ 39 m (128 ft) for Solution 3-C], CPCI 900 and 1200 sections are competitive because their unit costs are almost equal. In such cases, a final selection may be based on some alternative objective function (minimum weight or minimum reinforcement) or response criterion (minimum deflection).

Solution Sensitivity Considerations

The optimal solutions presented in this study are based on a set of parameters consistent with current design practice. The parameters to be considered in concrete bridge design are longitudinal configuration, bridge length and width, traf­fic loading, slab thickness, girder (standard or non-standard) section type, concrete grade, superstructure and substructure cost parameters.

In this section, the validity and possible application of the optimal solutions obtained outside the range of adopted pa­rameters are examined.

(a) Longitudinal Configuration - Optimal bridge sys­tems of 1-S, 2-S, 3-S, 2-C and 3-C type (Fig. 2) have been identified. The optimization procedure can be applied to any number of simple or continuous spans. For any span length, results of the present study of 2-C and 3-C configu­rations are still valid and conservative for 4-C, 5-C, etc., configurations.

(b) Bridge Length (L) -With the three standard girder sections and five longitudinal configurations considered, op­timal solutions for bridges not exceeding 100 m (328 ft) have been determined. Longer bridges may be designed with other than adopted standard girders (e.g., CPCI 1900, 2300 or AASHTO-PCI Type V, VI, etc.) or longitudinal configurations (4-C, 5-C, etc.). It also must be noted that for bridges with spans in excess of 40 m (131 ft), other struc­tural systems than reinforced concrete slab on precast, pre-

July-August 1993

stressed concrete !-girders may represent competitive alter­natives (prestressed concrete box girders or other forms of construction).

(c) Bridge Width (W)- Although optimization has been conducted for bridge widths W = 8, 12 and 16 m (26, 39 and 52 ft) , additional data are provided in Table 4 which lists the optimal number of girders for bridge widths varying between 8 and 16 m (26 and 52 ft). Determination of the prestressing force is possible by interpolating in Fig. 8 or by using Eq. (14).

(d) Traffic Loading- The study is based on the OHBD Code24 loading specifications for Class A highways (average daily truck traffic higher than 1000 or traffic higher than 4000).

In the new edition of the OHBD Code/ 5 the two 140 kN (31.5 kips) axles have been increased to 160 kN (36 kips), which will induce an increase in the live load moment of smaller than 14 percent. However, the corresponding impact factor (function of the number of axles governing the de­sign) is equal to 0.25 for the span lengths investigated in this paper. Since an impact factor of 0.4 has been assumed in this study, which is 60 percent higher than the new 0.25 value, these optimal solutions remain feasible within the new OHBD Code requirements.

It will also be shown in the next section that the design re­sults are comparable to those based on AASHTO HS20 loading/ 7 and may also be used for lighter loadings (e.g., AASHTO HS15 and H20 loadings).

(e) Slab Thickness (t) - Optimal solutions assume a slab thickness t = 225 mm (8 .9 in.), the minimum required by OHBDC. Optimal designs will remain valid for slab thick­nesses of at least 200 mm (8 in.) .

(f) Girder Section Type - The optimization program can readily include section variables in order to determine the optimal girder sections (Fig. 5a)22 in addition to system optimization variables. Consideration of general economy (formwork, fabrication and transportation costs) suggests that using standardized girder sections is more economical, although not structurally optimal. This is why the CPCI 900, 1200 and 1400 girders have been used in the present investi­gation. The CPCI 1900 and 2300 girders are not included because related costs are unavailable.*

(g) Concrete Grade (j~) - Girder optimization has as­sumed J; = 40 MPa (5800 psi). Results remain valid on the conservative side for higher concrete grades.

(h) Superstructure and Substructure Costs - Opti­mization of precast, prestressed concrete girder bridges has been investigated for the minimum cost of superstructure, assuming relatively moderate or low substructure costs. The effect of substructure costs are briefly discussed in the next section.

Although absolute costs of bridge systems will vary with the specific unit costs, the optimal design solutions will be insensitive to variations of these unit costs as long as their relative magnitudes remain unchanged.

* Private communication with Kris Bassi, Ministry of Transponation, Ontario, Canada, September 1990.

71

Page 13: Optimization of Precast Prestressed Concrete … of Precast Prestressed Concrete Bridge Girder Systems Z. Lounis, Ph.D. Research Assistant Department of Civil Engineering University

(' .

Table 5. Comparison of CPCI 23 and AASHTO-PCI2•25 precast concrete girder sections.

Moment of Top section Bottom section Girder types Depth,H Area, A inertia,/g modulus,S1 modulus,Sb and ratios (rnrn) (10' rnrn') (10' rnrn') (10' rnrn') (10' rnrn')

CPCI 900 900 218 19.30 38.40 48.60 AASHTO-PCI II 900 238 21.20 41.43 52.77

Ratio 1.000 0.916 0.910 0.927 0.921

CPCI 1200 1200 320 53.9 80.00 102.30 AASHTO-PCI III 1143 361 52.2 83.08 101.37

Ratio 1.050 0.886 1.032 0.960 1.010

CPCI1400 1400 413 102.6 134.2 161.4 AASHTO-PCI IV 1372 509 108.5 146.0 172.8

Ratio 1.020 0.811 0.946 0.920 0.934

Metric (SI) conversion factors: I mm = 0.0394 in.; I mm' = 0.00155 in.' ; I mm' = 0.000061 in.'; I mm• = 0.0000024 in.'

PRACTICAL RECOMMENDATIONS

Optimality Criterion for Prestressing Force

The service tensile stress constraint, which is always ac­tive for this type of bridge, may be adopted as an optimality criterion for the determination of the prestressing force:

P = [fst + (M8 + M5)/ Sb + (Ma + 0.75ML)/ Sbc] I (l/A + ejSb) (14)

where sb and sbc are the section moduli with regard to bot­tom fibers of the girder and composite (girder plus slab) sec­tions, respectively.

Substructure Cost Consideration

Fig. 10 shows that the optimization of the superstructure yields unique optimum bridge designs, except at two spe­cific length ranges where two configurations are competitive and the final choice is to be made by comparing their total (superstructure and substructure) costs. • For 27 < L s;; 32 m (89 < L s;; 105 ft), the superstructure

cost for the 1-S configuration (feasible for the CPCI 1400 girder only) is on average 24 percent (rather high) more expensive than the two-span continuous (2-C) configura­tion, which may be considered as the optimum (Fig. 10). However, if the cost of the intermediate pier in the 2-C configuration is very high (e.g., very difficult access site), the 1-S configuration may be preferable.

• For 44 < L s;; 55 m (144 < L s;; 180 ft), the 2-C and 3-C configurations may be competitive, and the. final choice will depend on their substructure cost ratio and the super­structure to substructure costs ratio, as is shown in the fol­lowing. Let 81 be the ratio of the total bridge costs for the 2-C and 3-C configurations:

81 = (sup+ sub )2_c

(sup+ sub )3_c (15a)

where "sup" and "sub" are the costs of the superstructure and substructure, respectively. Division of both numerator and denominator by sup2_c yields:

72

TypeD Type m Type IV

Fig. 11. AASHTO-PCI standard girder sections2•26 Types II, Ill

and IV. All dimensions in inches; 1 in.= 25.4 mm.

1 sub2_c +---

8 - SUP2-C

I - sup3-c sub3-c ---+---SUP2-C SUP2-C

(15b)

It is noted from Fig. 10 that the superstructure cost ratio sup3_c/sup2_c = 0.93 (constant) for all three bridge widths considered. Hence, the ratio 81 depends on two parameters:

~ = sub2_c /sup2_c (15c) and

~ = sub3_c/ sub2_c (15d)

8 - 1 + 82 l-

0.93+8283 (15e)

From Eq. (15e), it can be concluded that: • 3-C is optimal for ~"' 1. • 2-C is optimal for~> 1.15. • 3-C and 2-C are competitive for~< 1.15 and~ s;; 0.7.

AASHTO Implementation

(a) AASHTO-PCP·25 Girders vs. CPCP3 Girders -From Figs. 4 and 11, it is seen that the precast CPCI girders 900, 1200 and 1400 have very similar geometries with the AASHTO-PCI girders Types II, III and IV, respectively.

PCI JOURNAL

Page 14: Optimization of Precast Prestressed Concrete … of Precast Prestressed Concrete Bridge Girder Systems Z. Lounis, Ph.D. Research Assistant Department of Civil Engineering University

Table 6. Comparison of OHBDC24 and AASHT027 HS20 live load moments.

Code Simple span (1-S)

moments Span section and ratios 10 20 30 10

OHBDC 658 1872 3426 526

(kN-m)

AASHTO 439 1229 2018 332

(kN-m)

Rl 1.499 1.523 1.698 1.584

R2 0.941 0.956 1.066 0.994

Metric (SI) conversion factor: I kN-m = 8.850 kip-in.

Table 5, which compares the geometrical and mechanical characteristics of the two sets of girders, indicates that they have virtually identical section moduli S1 and Sb. Differences between CPCI and AASHTO-PCI section moduli are infe­rior to 8 percent for Types II and IV vs. CPCI 900 and 1400, and less than 4 percent for Type III vs. CPCI 1200. This sug­gests that the AASHTO-PCI Types II, III and IV are struc­turally equivalent to CPCI 900, 1200 and 1400, respectively.

(b) AASHT026 vs. OHBDC24 Design Requirements-

SLS Constraints

Allowable Stresses: OHBDC24 and AASHT026 adopt the same allowable tensile and compressive stresses at SLS, i.e., -0.5 .fJZ and OAf~, respectively.

Effective Service Moments:

OHBDC: M=MD+0.75 ML (16a)

AASHTO: (16b)

where MD and ML are the moments due to dead and live loads, respectively.

Although the OHBDC truck (lane) load moments are higher than the AASHTO (HS20) truck (lane) load moments (see Fig. 12), the OHBD Code considers only 75 percent of these moments at SLS. Moreover, transverse distribution factors in the OHBDC are larger than the AASHTO values (Dd = 1.8 to 2.15 vs. 1.67). Hence, OHBDC and AASHTO live load moments should be compared on the basis of their effective truck (or lane) load design moments, which in­clude the transverse distribution factor S/Dd, SLS load com­bination factor and impact factor / . These effective design live load moments M~ may be expressed as:

OHBDC: M~ = 0.75 (1 + /) (SID d) Mr~/2 (16c)

AASHTO: M~ = (1 + l) (S/1.67) M 1;/2 (16d)

where Mtll is the moment due to the governing truck or lane load. In the OHBD Code, 24 Dd is not a constant as in AASHTO, but a function of the number of lanes, torsional and flexural parameters of the bridge. In the new OHBD Code/ 6 the determination of Dd has been simplified to de­pend on the span length only (for a given bridge type). For

July-August 1993

Two-span continuous (2-C)

Span section Support section

20 30 10 20 30

1500 2892 468 1298 2394

970 1676 290 775 1512

1.546 1.720 1.614 1.675 1.583

0.970 1.080 1.013 1.052 0.994

comparison purposes, one may conservatively adopt Dd = 1.9. The ratio of OHBDC to AASHTO (HS20) maximum truck (lane) load moments, R1, and the ratio of their corre­sponding effective design moments, R2, are related by:

(16e)

For the simple and two-span continuous girders, Table 6 shows the maximum truck (lane) load moments by the two codes, their ratios R1 and ratios R2 of the effective design moments. The real differences between the AASHTO and OHBDC design service live load moments, illustrated by ratio R2, remain inferior to 6 percent, and, therefore, for all practical purposes, identical designs will satisfy SLS re­quirements of both codes.

ULS Constraints

The ultimate limit state (ULS) flexural requirements in OHBDC and AASHTO codes are:

OHBDC: Mu = 1.1 Mg + 1.2 Ms + 1.5 Ma + 1.4 ML :S: 0.85 M11 (5)

AASHTO:

In AASHTO, a constant overload factor is applied on the total dead load, while in OHBDC this factor varies from 1.1 (for girder weight moment) to 1.5 (for asphalt pavement weight moment). From Table 6, it can be assumed on aver­age that the OHBDC live load moment is 60 percent higher than the AASHTO moment R1 = 1.6. Taking the AASHTO live load moment as a reference, and including the trans­verse distribution factors, the OHBDC live load moment is 40 percent higher than the AASHTO moment (1.6 x 1.67/1.9 = 1.4). In addition, since the main dead load mo­ment is due to the girder and slab weight moments, the vari­ous dead load factors in Eq. (5) are substituted by a single factor, approximately 1.2, and the OHBDC ULS require­ment becomes:

In the new OHBD Code,25 the single understrength factor (0.85) has been substituted by strength reduction factors for concrete (0.75), steel (0.90) and prestressing steel (0.95).

73

Page 15: Optimization of Precast Prestressed Concrete … of Precast Prestressed Concrete Bridge Girder Systems Z. Lounis, Ph.D. Research Assistant Department of Civil Engineering University

60 140 !«) 200 160 oale load kN

30 70 70 100 80 wlleellood kN

~'~~+~i~~~~--'~~~* 1:

3.6. ~21 • 6.0 ·'· 7.2 L ~·-~----~~~~·xO ________ ~: ,, (typ.)

I

~

~ ! ms m ~m\ ~r-

J!:avel e = ~

"'[ e~Oil ~-ffi-iit-~~-o-a4:~! i~j

~ I .. t_ .

Q.25~

mm m r .. l~.- Plan 0.25

OHBDC truck load

H$20-44 fMS181 1000 lb 138 kNI 32JIIIO ft.t 1144 kNI H$110 .. fMS13.S 1000 lb 127 kNI :MJIIIO lb fUJI kNI

I 32JIIIO 1b 1144 kNI :MJIIIO 1b 1108 kNI

W • Combined -ight on tht lim two uloa wllid> io tht ,._ • for 1M---'"" H fM! trudt . V • Var-spoo:int-14 It to 30ft 14.2&7108.144 ml indlllift.S.Ocinvto•-

10 ft f3.0.S ml Clooronooand ---T mil~"' k~'j r~on

2ttl I Itt I 12ttt 111.110 "'' 11.130 .. , 11.110 "''

AASHfO HS ttuck load

Cllle L.Dad kN 412 9p98 140 112

f1~ II !Tfl ~I I I II II 'I 1 I II I II I' ltjq

1:3.6.~21• 6t~.o·'· 1.2 :1 1 Hiqhway Class A B Ct or Cz

Uniform!~ Distributed Lood 10 9 8 q =(kN/m)

(AI dhntnsions 111 •)

OHBDC lane load

IIIJIIIO .. ., kNIIaf­j...-~ 111111 ""1 2UDO. 1111 kNIIaf-~ / UnifO<m laldl40 pff 18.4 kN/ml of laid lono

AASHTO HS lane load

Fig. 12. OHBDC24 and AASHT027 live loadings: (a) OHBDC truck and lane loadings [Axles 2 and 3 have been changed to 160 kN (36 kips) in the new OHBDC25

] , (b) AASHTO HS20 truck and lane loadings.

74 PCI JOURNAL

Page 16: Optimization of Precast Prestressed Concrete … of Precast Prestressed Concrete Bridge Girder Systems Z. Lounis, Ph.D. Research Assistant Department of Civil Engineering University

However, since the optimal girder designs obtained are al­ways under-reinforced with low reinforcement indices, the corresponding equivalent understrength factor will be higher than 0.85. Hence, the original assumption (cf> = 0.85) is more conservative than the new code safety requirements.

Let A0 be an overall safety factor defined as the ratio of the ultimate resisting moment to the service moment. The overall safety factors A0 for the two codes are:

OHBDC: A0 =M/lf>MD+L = 1.41 + 0.953 ML/ MD+L (17c)

AASHTO: A0 = M/cf> MD+L = 1.44 + 0.967 ML/ MD+L (17d)

where Mv+L is the service dead and live load moment. The difference between the two safety factors does not exceed 2 percent for all practical values of MLIMD+V and it is con­cluded that the same design will be feasible for both the AASHTO and OHBDC codes.

Optimal Design Procedure

With the design aids resulting from the optimization study of precast, prestressed !-girder systems, the preliminary op­timal design of a bridge system consists of four simple steps to determine: (a) system selection, (b) optimum prestressing force and tendon layout, (c) non-prestressed steel for ULS flexural strength (if needed), and (d) bottom and top rein­forcements at piers.

(a) System Selection- Given the total bridge length L, width W and OHBDC or AASHTO (HS20) loadings, use Fig. 10, Table 2 and Table 4 to determine the optimal con­figuration, number and type of girders.

(b) Optimum Prestressing and Tendon Layout­Given the span length £, the optimum prestressing is ob­tained from Figs. 8(a) and 8(b) for simple and continuous span bridges, respectively. For bridge widths between the values investigated in this paper, interpolate linearly from Fig. 8 or use Eq. (14).

The tendon layout adopted for the type of girders investi­gated has two depressed points at the third span from each support (Fig. 5b ). The eccentricity at the middle third span (ec) should be taken as the maximum feasible value (i.e., for minimum concrete cover), while the eccentricity at the sup­port (ee) could have any value that satisfies the transfer stress constraints. Let "Ptim be a certain limit value for the ef­fective prestressing force:

where f3 is the ratio of the initial to effective prestressing forces. Thus:

(18b)

(18c)

(c) Non-Prestressed Steel for ULS Flexural Strength - Determine m from Fig. 9 as a function of the span length. Find the non-prestressed reinforcement from Eq. (7).

July-August 1993

(d) Bottom and Top Reinforcements at Piers- Deter­mine the bottom and top reinforcements at piers from Eqs. (13d) and (13e), respectively.

Of course, a final design and thorough check will follow and possibly result in appropriate adjustments of the opti­mum preliminary design.

Design Example 1: Member vs. System Optimization

The relevance and benefits of system optimization vs. member optimization are illustrated by a cost comparison of optimal designs for a reinforced concrete slab on a pre­cast, prestressed !-girder bridge with a total length L = 60 m (197 ft) and width W = 16 m (52 ft) (four lanes) for two cases: (a) specified longitudinal configuration and (b) vari­able longitudinal configuration.

(a) Design 1: Specified Configuration • Consider a two-span simply supported girder bridge (2-S)

with £ = 30 m (98 ft). • Optimum girder section is CPCI 1400, the only feasible

alternative for £ = 30 m (98 ft). • Design requires seven CPCI 1400 girders'' with P = 3700

kN (832 kips) and ec = 535 mm (21 in.) (maximum feasi­ble eccentricity).

• From Table 3, unit superstructure cost per deck area Z = $451 per m' (Z = $42 per sq ft), yielding a total super­structure cost of about $433,000. (b) Design 2: System Optimization

• From Fig. 10, the optimum solution for L = 60 m (197ft) and W = 16 m (52 ft) is the three-span continuous (3-C) bridge system.

• Optimum solution requires five CPCI 1200 girders. • From Fig. 8b, the optimum prestressing force is P = 2300

kN (517 kips) and ec = 427 mm (17 in.) (maximum feasi­ble eccentricity).

• From Fig. 10 or Table 3, unit superstructure cost per deck area Z = $335 per m' (Z = $31 per sq ft), yielding a total superstructure cost of about $322,000. The 25.7 percent cost reduction of Design 2 vs. Design 1

demonstrates the economic importance of system optimiza­tion vs. member optimization.

Design Example 2: AASHTO-PCI Application

A reinforced concrete slab on precast, prestressed AASHTO-PCI girders bridge of total length L = 130 ft (39.7 m) and width W =40ft (12.2 m) is to be designed ac­cording to AASHTO specifications for the HS20 loading. The slab thickness is 8.5 in. (216 mm) with an asphalt pavement of 3.5 in. (89 mm). For the girders, f~ = 6000 psi (41 MPa) and f~; = 4500 psi (31 MPa), and for the slab, f~ = 4500 psi (31 MPa). Also, fpu = 270 ksi (1860 MPa) and fy = 60 ksi (414 MPa).

(a) Optimal Bridge System Selection From Fig. 10, for L =130ft("' 40 m) and W =40ft ("'12m),

the best structural system consists of four CPCI 1200 two­span continuous [span length £ = 65 ft (20 m)] girders,

75

Page 17: Optimization of Precast Prestressed Concrete … of Precast Prestressed Concrete Bridge Girder Systems Z. Lounis, Ph.D. Research Assistant Department of Civil Engineering University

equivalent to four AASHTO-PCI Type III girders and S = 9.5 ft (2.9 m) girder spacing.

(b) Optimum Prestressing and Tendon Layout From Fig. 8b and for a span length of 65 ft (""20 m),

the optimum prestressing for the AASHTO-PCI Type III is 483 kips (2150 kN) with eccentricity at midspan of 17 in. ( 427 mm). The tendon layout is depressed at the third span points [21.6 ft (6.58 m)] from each support, as shown in Fig. 5b. The end eccentricity is easily determined from the trans-fer stress constraints. '

(c) Non-Prestressed Steel for ULS Flexural Strength From Fig. 9b, for £ = 65 ft (20 m) (and AASHTO-PCI

Type III), the total reinforcement index is m = 0.03. The amount of non-prestressed reinforcement is obtained from Eq. (7):

As= (Wbd f~ -Aps fps)lfy = 0.37 in.2 (239 mm2)

(d) Bottom and Top Reinforcements at Piers Eqs. (17d) and (17e) yield the positive (bottom) and nega­

tive (top) moment reinforcements at the piers as 3.2 in. 2

(2065 mm2) and 6.9 in.2 (4452 mm2

), respectively.

CONCLUSIONS

1. The general optimization approach presented in the paper is ideally suited for determining the optimal designs of precast, prestressed concrete sections, members or bridge systems.

2. Standardization of optimal bridge systems through opti­mization of longitudinal and transverse bridge configura­tions is the main objective of the paper. The overall design economy may rationally be improved by optimizing bridge systems rather than their (standardized) components.

3. Optimal longitudinal configurations are simply sup­ported girders (1-S) for bridge lengths not exceeding 27m (89 ft); two-span continuous girders (2-C) for bridge lengths varying between 28 and 44 m (92 and 144 ft); and three­span continuous girders (3-C) for bridge lengths between 55 and 100m (180 and 328ft). For bridge lengths of 44 to 55 m (144 to 180ft), both optimal superstructure systems 3-C and 2-C are competitive as long as the substructure cost of the former remains moderate; otherwise, the two-span continu­ous system is more economic.

4. Optimal girder sections: For total bridge lengths L :::; 66 m (217 m), the optimal system always consists of CPCI 1200 sections [with three, four or five girders for W = 8, 12 or 16m (26, 39 or 52ft), respectively]. For L > 66 m (217 ft), the optimal system always consists of CPCI 1400 sections (with four to seven girders depending on Wand L) (Fig. 10 and Table 4).

5. Optimal system selection may be obtained directly from Fig. 10 by specifying the bridge length (L) and width (W). Although Fig. 10 identifies optimal bridge systems under OHBDC traffic loading, it may conservatively be used for lighter loading schemes.

6. Maximum feasible girder spacing is insensitive to bridge width and depends only on the bridge configuration (simple or continuous), span length, girder type and slab

76

thickness. However, the optimal girder spacing depends on the bridge width in addition to all other parameters.

7. Prestressing reinforcement of !-girders may be found directly from Fig. 8, which corresponds to maximum feasi­ble eccentricity (i.e., minimum concrete cover) at midspan. The support eccentricity may be determined from Eqs. (18b) and (18c). Additional non-prestressed reinforcement re­quired at midspan may be determined from Fig. 9 and Eq. (7). The non-prestressed continuity reinforcement at the piers is determined from Eqs. (13d) and (13e).

The total reinforcement index of precast girders is rather low and is marginally affected by the bridge width. It in­creases with span length and decreases with girder depth.

8. Although the superstructure costs of various bridge sys­tems (Table 3) depend on the adopted unit costs, optimal bridge systems are unaffected by absolute values of the lat­ter as long as their ratios remain comparable. Hence, the system effectiveness will conserve the trends found in this investigation, while real costs may vary with the specified economic conditions of a particular project.

9. Design aids provided in the paper (Figs. 8, 9 and 10 and Tables 2 and 4) may be used to generate preliminary designs, to be detailed in compliance with any additional economic or structural requirements (shear strength, deflec­tion and other parameters). Alternatively, basic design pa­rameters may be determined by using the prestressing force optimality criterion.

10. Sensitivity considerations of optimal solutions suggest that results obtained in this study are applicable to most practical values of design parameters. Also, these results are representative of general trends in optimizing bridge config­urations and may be used for preliminary designs with bridge standard specifications other than OHBDC.

11. AASHTO implementation of bridge system optimiza­tion is direct and only requires substitution of standard sec­tions AASHTO-PCI Type II, III and IV for CPCI 900, 1200 and 1400 sections, respectively.

12. Extensions of bridge system optimization procedures are necessary and will be considered separately. In particu­lar, the following topics deserve further investigation: opti­mization of precast, prestressed concrete !-girder sections to be standardized, other competitive transverse configurations (prestressed concrete voided slab or box girder, composite steel girder-concrete deck, etc.) and comprehensive sensitiv­ity studies.

ACKNOWLEDGMENT The work reported in this paper is part of a Ph.D. thesis

prepared by the first author under the direction of the second author, in partial fulfillment of the degree requirements in the department of civil engineering, University of Water­loo, Waterloo, Ontario, Canada.

The financial support of the Natural Sciences and Engi­neering Research Council (NSERC) of Canada under Grant A-4789 is gratefully acknowledged. The authors are in­debted to the Ontario Ministry of Transportation for provid­ing valuable bridge design data and to the PCI JOURNAL reviewers for their constructive comments.

PCI JOURNAL

Page 18: Optimization of Precast Prestressed Concrete … of Precast Prestressed Concrete Bridge Girder Systems Z. Lounis, Ph.D. Research Assistant Department of Civil Engineering University

REFERENCES 1. ACI-ASCE Committee 343, "Analysis and Design of Rein­

forced Concrete Bridge Structures," ACI Manual of Concrete Practice, Part 4, American Concrete Institute, Detroit, MI, 1989.

2. Precast Prestressed Concrete Short Span Bridges - Spans to 100 feet, Second Edition, Precast/Prestressed Concrete Insti­tute, Chicago, IL, 1981.

3. Heins, C. P., and Lawrie, R. A., Design of Modem Concrete Highway Bridges, John Wiley & Sons, New York, NY, 1984.

4. Dunker, K. F., and Rabbat, B. G., "Performance of Prestressed Concrete Highway Bridges in the United States - the First 40 Years," PCI JOURNAL, V. 37, No. 3, May-June 1992, pp. 48-64.

5. Kirsch, U., "Optimum Design of Prestressed Plates," ASCE Journal of the Structural Division, V. 99, No. ST6, June 1973, pp. 1075-1090.

6. Wills, J., "A Mathematical Optimization Procedure and Its Applications to the Design of Bridge Structures," Transport and Road Research Laboratory, Report LR 555, Crowtorne, Berkshire, England, 1973, pp. 1-28.

7. Mafi, M., and West, H. H., "Cost-Effective Short Span Bridge Systems: A Selection Concept and an Optimization Procedure, Second International Conference on Short and Medium Span Bridges, V. 2, Ottawa, Ontario, Canada, 1986, pp. 117-131.

8. Naaman, Antoine, "Computer Program for Selection and Design of Simple-Span Prestressed Concrete Highway Girders," PCI JOURNAL, V.l7,No.l,January-February 1972,pp. 73-81.

9. Torres, G. Guzman Baron, Brotchie, J. F., and Cornell, C. A., "A Program for the Optimum Design of Prestressed Concrete Highway Bridges," PCI JOURNAL, V. 11, No. 3, June 1966, pp. 63-71.

10. Rabbat, B. G., and Russell, H. G., "Optimized Sections for Pretensioned Concrete Bridge Girders in the United States," First International Conference on Short and Medium Span Bridges, V. 2, Toronto, Ontario, Canada, 1982, pp. 97-111.

11. Cohn, M. Z., and Mac Rae, A. J., "Prestressing Optimization and Design Practice," PCI JOURNAL, V. 29, No. 4, July­August 1984, pp. 68-83.

12. Templeman, A. B., and Winterbottom, S. K., "Optimum De­sign of Concrete Cellular Spine Beam Bridge Decks," Pro­ceedings, Institution of Civil Engineers, Part 2, V. 59, 1975, pp. 669-697.

13. Aguilar, R. J., Movassaghi, K., and Brewer, J. A., "Computer­ized Optimization of Bridge Structures," Computers & Struc­tures, V. 3, 1973, pp. 429-442.

14. Bond, D., "An Examination of the Automated Design of Pre­stressed Concrete Bridge Decks by Computer," Proceedings, Institution of Civil Engineers, Part 2, V. 59, 1975, pp. 669-697.

15. Bond, D., "Optimal Design of Continuous Prestressed Con­crete Bridges," International Symposium on Nonlinearity and Continuity in Prestressed Concrete (M. Z. Cohn, editor), V. 3, University of Waterloo, Waterloo, Ontario, Canada, 1983, pp. 201-234.

16. Aziz, E. M., and Edwards, A. D., "Some Aspects of the Economics of Continuous Prestressed Concrete Bridge Gird­ers," The Structural Engineer, V. 44, No. 2, February 1966, pp. 49-54.

17. Adin, M., Cohn, M. Z., and Pinto, M., "Comprehensive Opti­mal Limit Design of Reinforced Concrete Bridges," Inter-

July-August 1993

national Symposium on Nonlinear Design of Concrete Struc­tures, University of Waterloo, Waterloo, Ontario, Canada, 1979, pp. 349-377.

18. McDermott, J. F., Abrams, J. I., and Cohn, M. Z., "Some Re­sults in the Optimization of Tall Building Systems," IABSE Ninth Congress - Preliminary Report, Amsterdam, Nether­lands, 1972, pp. 855-861.

19. McDermott, J. F., Abrams, J. I., and Cohn, M. Z., "Computer Program for Selecting Structural Systems," ASCE Annual Meeting, Houston, TX, Preprint 1863, October 1972.

20. Gellatly, R. A., and Dupree, D. M., "Examples of Computer­Aided Optimal Design of Structures," IABSE Tenth Congress -Introductory Report, Tokyo, 1976, pp. 77-105.

21. Kulka, F., and Lin, T.Y., "Comparative Study of Medium­Span Box Girder Bridges With Other Precast Systems," First International Conference on Short and Medium Span Bridges, V. 1, Toronto, Ontario, Canada, 1982, pp. 81-94.

22. Lounis, Z., and Cohn, M. Z., "Optimal Design of Prestressed Concrete Highway Bridge Girders," Proceedings, Third Inter­national Symposium on Concrete Bridge Design, Washington, D.C., March 1992 (to be published).

23. CPCI, Metric Design Manual, Second Edition, Canadian Pre­stressed Concrete Institute, Ottawa, Ontario, Canada, 1987.

24. OHBDC-1983, Ontario Highway Bridge Design Code and Commentary, Second Edition, Ministry of Transportation and Communications, Downsview, Ontario, Canada, 1983.

25. OHBDC-1992, Ontario Highway Bridge Design Code Draft, Third Edition, Ministry of Transportation and Communica­tions, Downsview, Ontario, Canada, 1992.

26. PC/ Design Handbook - Precast and Prestressed Concrete, Fourth Edition, Precast/Prestressed Concrete Institute, Chicago, IL, 1992.

27. AASHTO, Standard Specifications for Highway Bridges, Fif­teenth Edition, American Association of State Highway and Transportation Officials, Washington, D.C., 1992.

28. Geren, K. L., Abdelkarim, A. M., and Tadros, M. K., "Precast/ Prestressed Concrete Bridge !-Girders: The Next Generation," Concrete International, V. 14, No.6, June 1992, pp. 25-28.

29. Bakht, B., Cheung, M. S., and Aziz, T. S., "Application of a Simplified Method of Calculating Longitudinal Moments to the Ontario Bridge Design Code," Canadian Journal of Civil Engineering, March 1979, pp. 36-50.

30. Bakht, B., and Jaeger, L. G., Bridge Analysis Simplified, McGraw-Hill Book Company, New York, NY, 1985.

31. Neville, A. M., Dilger, W. H., and Brooks, J. J., Creep of Plain and Structural Concrete, Construction Press, 1983.

32. Freyermuth, C. L., "Design of Continuous Highway Bridges With Precast, Prestressed Concrete Girders," PCI JOURNAL, V. 14, No.2, Aprill969, pp. 14-39.

33. ACI Committee 209, "Prediction of Creep, Shrinkage and Temperature Effects in Concrete Structures," ACI Manual of Concrete Practice, Part 1, American Concrete Institute, De­troit, MI, 1989.

341 Murtagh, B. A., and Saunders, M.A., "MINOS/Augmented," User's Manual, Department of Operations Research, Stanford University, Stanford, CA, Technical Report SOL 80-9, 1980.

35. Brook, A., Kendruck, D., and Meeraus, A., "GAMS- Gen­eral Algebraic Modeling System," User's Guide, Scientific Press, 1988.

77

Page 19: Optimization of Precast Prestressed Concrete … of Precast Prestressed Concrete Bridge Girder Systems Z. Lounis, Ph.D. Research Assistant Department of Civil Engineering University

APPENDIX A- NOTATION

The following symbols are used in this paper in addition to the standard ACI and PCI notation:

A = girder cross-sectional area Ccs = concrete cost in slab and diaphragms cdj = cost of deck joints Cg = cost of precast girder per length

Cpc = unit cost of positive moment connection at piers

c. = non-prestressed steel unit cost cbot = bottom concrete cover crop = top concrete cover D d = design load distribution factor

Dv Dy = bridge flexural rigidity in transverse and longi­tudinal directions, respectively

Dxy, Dyx = bridge torsional rigidity in transverse and lon­gitudinal directions, respectively

D1, D2 = bridge coupling rigidity in transverse and lon­gitudinal directions, respectively

fsP fsc = allowable tensile and compressive stresses at service, respectively

ftp frc = allowable tensile and compressive stresses at transfer, respectively

ee, ec = tendon eccentricities at end and midspan sec­tions, repectively

I = impact factor lg, lgc = moments of inertia of girder and composite

sections, respectively L = total bridge length c = span length

£* = equivalent span length for live load moment computation in continuous girders

M, Mv+L = total service moment

78

Ma = asphalt pavement weight moment Mcsn = moment at support n due to combined effects

of creep and shrinkage Mvn = support n restraint moment due to dead load

Mg = girder weight moment ML = live load moment Mn = nominal resisting moment; moment at support

n if structure is continuous and monolithically cast in a single stage

Mpn = support n restraint moment due to prestressing M8 = slab weight moment

Mshn = support n restraint moment due to shrinkage Mu = ultimate load moment M~ = moment induced by creep at support n

n = number of girders nP = number of piers

n. = number of critical span sections R1 = ratio of OHBDC and AASHTO HS20 truck

(lane) load moments R2 = ratio of OHBDC and AASHTO HS20 effective

design moments S = girder spacing S' = slab overhang

Sb, Sbc = section moduli with regard to bottom fibers of girder and composite sections

sup = superstructure cost sub = substructure cost

t = slab thickness t

0 = concrete age at loading

t1 = concrete age at which continuity is achieved Vcs = volume of concrete in slab and diaphragms Vpc = volume of positive moment reinforcement

W = bridge width w. = weight of non-prestressed steel in slab and

diaphragms Wsn = weight of negative moment steel in slab at

piers wsp = weight of positive moment steel in girders

y = distance between centroids of girder and com­posite sections

Ybc = distance from centroid of composite section to bottom fiber

y1 = distance from centroid of girder section to top fiber

Z = superstructure cost per deck area a = bridge torsional parameter f3 = ratio of initial and effective prestressing forces 81 = ratio of total bridge costs of 2-C to 3-C config-

urations 8z = ratio of substructure to superstructure costs for

2-C configuration ~ = ratio of substructure costs of 3-C to 2-C config-

urations ¢ = flexural strength reduction factor X = aging coefficient

.A.o = ratio of ultimate resisting moment to service moment

v (t, t0 ) = creep coefficient (ratio of creep strain to initial elastic strain)

vu = ultimate creep factor m = total reinforcement index

acP acs = concrete stresses at transfer and service, re­spectively

8 = bridge flexural parameter

PCI JOURNAL