Structural Redundancy of Prestressed Concrete Box Girder Bridges

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

Structural Redundancy of

Prestressed Concrete Box Girder Bridges

Jian Yang1* , O. Murat Hamutcuoglu2 , Yingsheng Ni3 , Michel Ghosn4

1
Bridge Engineer, Ph.D. candidate
*Corresponding author
McLaren Engineering Group
100 Snake Hill Road, West Nyack, NY 10994
Department of Civil Engineering,
The City College of New York / CUNY,
E- mail: jyang@mgmclaren.com
Tel. (845)353-6400 Fax. (845)353-6509
2
Structural Engineer, Ph.D., P.E.
HNTB Corporation
5 Penn Plaza, 6th Floor, New York, New York 10001
E- mail: ohamutcuoglu@hntb.com
Tel. (212) 594-9717 Fax (212) 947-4030
3
Ph.D. candidate
Department of Bridge Engineering,
Tongji University, Shanghai 200092, P.R. C hina
E- mail: niyingsheng2008@aliyun.com
Tel. +8618818260850
4
Professor of Civil Engineering,
The City College of New York / CUNY,
160 Convent Ave. New York, NY, 10031.
E- mail: mghosn@ccny.cuny.edu
Tel. (212)650-8002 Fax.(212)650-6965

Word Count
Text - 4250
Tables (3) -750
Figures (10) – 2500
Total = 7500 words
Final-submission Date: Nov.15, 2015
1 Abstract
2
3 Bridges are usually designed and evaluated based on member strength and serviceability criteria,
4 however, it is expected that they exhibit sufficient levels of reserve strength and multiple paths to
5 resist collapse should any of their members get damaged or exceed its nominal strength capacity.
6 The ability of a structural system to redistribute the load around damaged members is referred to
7 as structural redundancy. To account for bridge redundancy during the design and safety
8 evaluation process, the AASHTO LRFD and LRFR provide a preliminary set of load modifiers
9 or system factors most of which were based on the code writers judgment and experience.

10 This paper evaluates the redundancy of prestressed concrete box girder bridges under vertical
11 loads and lateral loads. Pushover and pushdown analyses are performed using frame and grillage
12 models in SAP2000. The grillage models are validated by more advanced finite element models
13 built in ABAQUS. The effects of different damage scenarios and types of connections between
14 the box girder and columns on the structural redundancy level are investigated. Based on those
15 results, a set of system factors for prestressed concrete box bridges accounting for the system
16 redundancy is proposed. The proposed system factors can be used during the design and safety
17 assessment of prestressed concrete box girder bridges subjected to transient lateral load and
18 vertical vehicle load.

19 Keywords: structural redundancy; system factor; nonlinear behavior; prestressed concrete box
20 girder.

1
21 Introduction
22
23 Traditionally, the design and load capacity evaluation of highway bridges have been executed on
24 a member by member basis. Yet, bridges are expected to have sufficient levels of structural
25 redundancy to sustain the failure of a main member and be still able to carry significant load to
26 allow for the evacuation of the structure and avoid loss of life before the damage is detected and
27 bridge closure or corrective actions are undertaken. However, the quantification of redundancy
28 is not fully formulated in current bridge standards and specifications. A 1985 “State of the Art
29 Report on Redundant Bridge Systems” concluded that although analytical techniques to study the
30 response of damaged and undamaged flexural systems to high loads are available, “little work
31 has been done on quantifying the degree of redundancy that is needed” (ASCE-AASHTO Task
32 Committee; 1985). Since that time, research studies and design guides and specifications have
33 proposed different approaches for evaluating the structural redundancy in b ridges. A
34 comprehensive review and synthesis of issues related to the redundancy ana lysis of bridges was
35 presented by Dexter et al (2005).
36
37 In a first attempt at providing a method to incorporate redundancy criteria in the bridge design
38 specifications, the AASHTO LRFD (2012) proposed the adoption of load modifiers in the design
39 check equations to account for redundancy during the design of new bridges based on the
40 recommendation of Frangopol and Nakib (1991). Specifically, the AASHTO LRFD
41 recommended applying different load modifiers varying between 0.95, 1.00 or 1.05 to reflect the
42 levels of bridge redundancy and ductility. An additional load modifier was related to the
43 importance of the structure for defense/security consideration. However, the specifications d id
44 not explain how to identify which bridges have low and high redundancy or how to define low
45 and high ductility. As explained in the AASHTO LRFD (2012) Commentary, the recommended
46 values had been subjectively assigned pending additional research. An alternate approach
47 adopted by the Canadian code CAN/CSA-S6-06 (2006) directed bridge engineers to use different
48 resistance factors to reflect different target member reliability index values which were selected
49 based on the redundancy of the bridge system expressed in terms of the consequence of a
50 member’s failure and the ductility of the member being evaluated. However, like the AASHTO
51 LRFD, this approach relied on the judgment of the engineer in deciding which bridges are
52 redundant and in judging the consequence of a member’s failure. The LRFR option of the
53 AASHTO Manual for Bridge Evaluation (MBE, 2011) assigned a system factor to be applied on
54 the resistance side of the equation with values ranging between 0.85 and 1.0 for bridge
55 configurations that have been demonstrated to have low levels of redundancy. A table provided
56 some guidelines as to how to assign the appropriate system factor based on bridge geometries
57 and configurations. Following the same line of thinking, some state load rating manuals such as
58 the Florida DOT (2012) had also developed their own sets of system factors. But, as stated by
59 Mertz (2006), these were primarily based on very limited analyses and heavily relied on
60 “engineering judgment”.
61
62 A series of studies to develop approaches and criteria for the quantification of bridge system
63 redundancy were undertaken under the auspices of the National Cooperative Highway Research
64 Program (NCHRP) by Ghosn and his colleagues (Ghosn and Moses, 1998; Liu, Ghosn, et al.,
65 2001; Yang and Ghosn, 2015). Many researchers/engineers (Hunley and Harik 2012; Hovell,
66 2007; Mertz, 2006), consulting companies (HNTB) and code writing organizations (AASHTO

2
67 LRFR; Florida DOT; Wisconsin DOT) have adopted the redundancy concepts developed in
68 NCHRP Reports 406 and458 and specified similar system factors as a way to account for
69 redundancy during the design or the safety evaluation of typical short and medium span bridge
70 systems. However, NCHRP 406 and 458 did not provide system factors for some bridge system
71 and subsystem configurations that have become more popular in recent years such as multi-cell
72 box girder bridges. Also, the two reports did not verify the applicability of the system factors
73 that they proposed for analyzing the entire bridge system including the interaction between the
74 superstructure and substructure.
75
76 This paper summarizes and validates the results of prestressed concrete box girder bridges
77 published in the appendices of NCHRP Report 776 (Ghosn & Yang, 2014) and proposes system
78 factors. As prestressed concrete box girder bridges offer a good option when building curved
79 bridges, they are becoming a popular choice for new designs due to their high torsion stiffness.
80 The proposed system factors will be of immediate interest to bridge design engineers. The
81 objective of this paper is to investigate the redundancy level of this type of bridges and direct
82 bridge engineers to use the proposed system factors to account for the structural redundancy
83 under transient lateral and vertical loads.

84 Bridge Redundancy
85
86 Figure 1 gives a conceptual representation of the behavior of a bridge structure and the different
87 levels that should be considered when evaluating member safety, system safety and system
88 redundancy. The load multipliers LF represent the capacity of the system in terms of the number
89 of design trucks that it can carry to reach different limit states. Specifically, LF u represents the
90 ability of the originally intact system to resist system collapse; LF f represents the functionality
91 limit; and LF1 corresponds to first member failure.

Load Factor

LF u Intact system
LF f

Assumed linear
behavior
LF 1

LF d Damaged bridge

First member Ultimate Loss of Ultimate Bridge Response


failure capacity of functionality capacity of
damaged system intact system
Figure 1. Representation of typical behavior of bridge systems

3
92 If redundancy is defined as the capability of a structure to continue to carry loads after the failure
93 of the most critical member, then comparisons between the load multipliers LF u, LFf, LFd and
94 LF1 would provide non-subjective and quantifiable measures of system redundancy. Thus, the
95 following three deterministic measures of system redundancy may be defined in terms of the
96 ratio of the system’s capacity as compared to the most critical member’s capacity:
97
LFU
Ru 
LF1
98
LFf
Rf 
LF1
99 Eq. (1)
LF
Rd  d
LF1
100
101
102 where Ru=system reserve ratio for the ultimate limit state, Rf=system reserve ratio for the
103 functionality limit state, Rd= system reserve ratio for the damage condition. Ru, Rf and Rd can
104 thus be used as measures of system redundancy as they represent the ability of a system to carry
105 load beyond the failure of the most critical member.

106 System factors


107
108 NCHRP Reports 406 and 776 proposed that structural system redundancy be considered during
109 the design/safety check process by applying system factors such that:

110 s RnN    iQi Eq. (2)

111 Where RnN is the required member capacity accounting for bridge redundancy,  s is the system
112 factor,is the member resistance factor as specified in the current AASHTO LRFD Bridge
113 Design Specifications, i is the load factor for load i, Q i is the load effect of load i.

114 The system factor,  s of Equation (2) provides a measure of the system reserve strength as it
115 relates to ductility, redundancy and operational importance, and their interaction. The system
116 factor,  s is related to two other factors,  su and sd which respectively account for system
117 functionality and resistance to collapse conditions after the strength of the most critical member
118 of an originally intact bridge is exceeded, and for the system strength of a bridge in damage state
119 condition. Functionality is defined as the ability of the originally intact system to carry heavy
120 live loads without exceeding vertical deflection limits equal to span length/100 which would
121 render the bridge unfit for use. Damaged conditions include failure of a main girder of multi-
122 girder bridges, the failure of an external web of one box of a box girder bridge or the loss of one
123 column of multi- column bents.
124
125 Non-redundant bridges are penalized by requiring their members to provide higher safety levels
126 than those of similar bridges with redundant configurations. The aim of  s, is to add reserve
127 capacity for non-redundant systems so that the overall system reliability is increased. If adequate

4
128 redundancy levels are present, a system factor  s=1.0 is used. In the instances where the level of
129 redundancy is more than adequate, a value of  s greater than 1.0 may be used. An upper limit
130 equal to 1.20 is recommended for  s until more experience is gained in the application of these
131 factors in actual design and safety check situations.

132 Bridge Description


133
134 A multi-cell prestressed concrete box girder bridge with two-column bents is shown in Figure 2.
135 The four-cell box girder deck has the top slab width of 58’ 10”. The three-span continuous
136 bridge is 412 ft long and the span lengths are 126 ft, 168 ft and 118 ft. The thicknesses of the top
137 and bottom slabs are 9 1/8” and 8 1/4”, respectively. The depth of the box girder is 6’ 9”. The
138 connection between the superstructures and the substructures are through the integral cap beams
139 with the same depth of the girders and 8 ft width. The origina l bridge superstructure is connected
140 to two 6 ft diameter round columns through the moment resisting rigid connections. In the
141 further analyses, the effect of the moment- free bearing configuration is also investigated.
142
143 The unconfined concrete strength is assumed to be 4000 psi and the reinforcing steel is taken as
144 Grade 60 with the ultimate stress capacity is 90 ksi. The prestressing tendons have the peak
145 stress of 270 ksi and their yield stress is 230 ksi. The area of tendons in each section is 9 in.2 and
146 the initial prestressing force is set as 1,824 kips without the losses. The total loss in the
147 prestressing force is assumed 20 ksi including the elastic shortening, creep, shrinkage and the
148 steel relaxation stresses. The tendons are considered in parabolic geometry along the girders and
149 all section moment-curvature calculations include the exact location of the tendons in the
150 sections analyses.

5
(a) Elevation view

(b) Typical section and connection between girder and columns

Figure 2. Bridge Configurations

151 Lateral Pushover Analyses


152 Model 1 with Rigid Cap-Column Connections

153 The lateral pushover analysis of the original model is performed using SAP2000. The plastic
154 hinge definitions are applied at the bottom and top of the columns to capture the nonlinear axial-
155 flexure interaction, as shown in Figure 3(a). The lateral pushover analysis is initiated with a
156 nonlinear dead load, tendon jacking and live load analysis that provide the initial stresses in the
157 elements. Figure 3(b) shows failure modes and plastic hinge propagation during the pushover
158 analysis. The dead load is calculated through the self- weight feature of SAP2000 and then the
159 prestressing tendon jacking forces are applied to provide the initial condition of the bridge prior
160 to an extreme event case.

6
161 Model 2 with Pinned Connections

162 Model 2 assumes that the cap beam is connected to the columns through bearings applied
163 between the top of the column and the cap beam, as illustrated in Figure 3(c). Elastic links are
164 used to model the pads that serve to release the rotations at the top of the columns with the
165 rotational stiffnesses of these bearings being set at zero. Failure modes and plastic hinge
166 propagation are shown in Figure 3(d).
167
168 The sectional moment curvature response of the column is illustrated in Figure 4 where the
169 extreme concrete fibers fail due to the strain exceeding the limit spalling strains and then the
170 compression fibers in the confined concrete region reach the ultimate compressive strain limits.
171 The bridge is loaded by 20% of the HL93 load along two lanes to represent the regular traffic
172 that may be on the bridge during an earthquake.
173

174 Figure 5 compares the results of the pushover analysis performed for Model 1 and Model 2.
175 According to linear elastic pushover analysis o f Model 1, the lateral force that causes the first
176 member to reach its load carrying capacity P1 is 501 kips when the bottom section of the pier
177 reaches its flexure capacity. Following a nonlinear pushover analysis, the ultimate capacity of the
178 system or collapse is reached when the total load Pu=720 kips. According to Equation (1), the
179 redundancy ratio of Model 1, Ru, is equal to 720 kips/ 501 kips=1.44. Similarly, the redundancy
180 ratio of Model 2, Ru, is equal to 345 kips/ 261 kips=1.32, which is 8.3% lower than that of Model
181 1.

7
(a) Model 1: Cross Sections with integral connection between cap beam and column

(b) Model 1: column top and bottom sections exceed elastic limits; concrete compressive
strains exceed crushing limit.

(c) Model 2: Cross sections with elastic links (pinned bearings) at the pier

(d) Model 2: column bottom sections exceed their elastic limits; concrete compressive
strains exceed crushing limit.

Figure 3. Transverse Pushover Analysis

8
Figure 4. Column Section Behavior Showing Compressed Concrete Region in Dark Blue and
Moment Curvature under the DL and 20% LL axial force.

Model 1 & 2 Pier Transverse Pushover


800
Pier Total Base Shear (kips)

700 Model 1: Rigid


600
500 Model 2: Bearing
400
300 Model 1: Rigid Linear
Analysis
200
Model 2: Bearing-
100 Linear Analysis
0
0 0.2 0.4 0.6 0.8 1
Pier Top Displacement (ft)

Figure 5. Transverse Pushover Analysis

9
182 Vertical Pushdown Analyses
183 Model 3 with SAP Grillage

184 The nonlinear pushdown analysis under the effect of vertical load is not as straightforward as the
185 lateral pushover analyses performed for Models 1 and 2. The nonlinear behavior of the box
186 girder can be simulated by plastic hinge theory if the multi-cell box girder is modeled using a
187 grillage where the transverse and longitudinal stiffnesses of the elements can be approximated.
188 The flexural and torsional stiffnesses of the grillage elements are calculated through existing
189 methodologies proven to capture box girder behavior accurately (Hambly, 1991; Zokaie, et al.,
190 1991). Two HS-20 trucks with 4 ft. lateral distance are applied at the mid-span, as shown in
191 Figure 6.

(a) 3D grillage model with two side by side HS-20 trucks

(b) Moment-Curvature Curve

Figure 6. Grillage Frame Model 3 with the HS-20 Trucks and Moment-Curvature Curve of
Grillage Segments

10
192 Comparison of SAP and ABAQUS Results

193 The purpose of this section is to validate the use of grillage models by SAP2000 for the analysis
194 of superstructures under the effect of vertical loads. Bridge superstructure is the same as Model
195 3, but use pinned supports between superstructures and substructures to simplify the modelling in
196 SAP and ABAQUS. Figure 7 shows a 3-D ABAQUS model, meshed section, prestressed
197 tendons layout, reinforcements and load positions of two HS-20 trucks. The vertical force of the
198 nonlinear pushdown analysis is normalized by the total weight of two HS-20 trucks, 144 kips.
199 Figure 8 shows the ultimate collapse load estimated in the grillage analysis is only 6.6% lower
200 than the ABAQUS result. The results confirm that the grillage model does an adequate job of
201 predicting the failure load.

11
(a) 3-D model (b) Section mesh

(c) Prestressed tendons (d) Reinforcements layout

(e) Two HS-20 truck loads


Figure 7. ABAQUS Finite Element Model

25
Number of Design Trucks

20

15

10
ABAQUS
5 SAP2000_Plastic hinge length=0.5 girder depth
SAP2000_Plastic hinge length=1.0 girder depth
0
0 1 2 3 4
Mid-Span Vertical Displacement (ft)
Figure 8. Comparison of Pushdown Results in SAP2000 and ABAQUS

12
202 Model 4 with Column loss

203 In Model 4, one column is removed from the system to study the capacity of the system to carry
204 some load following major damage to one of the columns. The damage scenario could represent
205 a situation where one of the column is hit by a truck, ship or debris carried by a flood or when
206 one column foundation is damaged to a major scour or if one of the columns has been exposed to
207 major deterioration or construction errors.
208
209 Two cases are considered. The first case is given in Model 4-a where the columns sections are
210 assumed to behave within linear elastic limits. The second case, Model 4-b, is a more realistic
211 approach that represents the nonlinear material behavior in both superstructures and
212 substructures. The numerical results of the vertical pushdown analyses on the Models 4-a and 4-
213 b are compared to the original case (Model 3), as shown in Figure 9. Regarding the results from
214 Model 4-a where the substructure columns is in elastic manner, the prestressed girders are
215 significantly redundant when one of the piers in the structure is subjected to a column loss.
216 Model 4-a experienced the girder failure at the negative moment section of the exterior girder at
217 load factor (aka, number of design trucks) LFd=23.8. The redundancy ratio of Model 4-a,
218 Rd=23.8/18.0=1.32. Following a nonlinear pushdown analysis, the ultimate capacity of the
219 system or collapse for Model 4-b is reached when the total load LFd=19.0. Therefore, the
220 redundancy ratio of Model 4-b, Rd =19.0/18.0 =1.06.

Model 4: Vertical Pushdown with Column Losses


30

25
Number of Design Trucks

20

15

10

0
0 1 2 3 4
Mid-Span Vertical Displacement (ft)
Model 3: Original Case Model 4-a: Column Loss
Model 4-b: Column Loss with Column Hinge Model 4 Linear
Model 4: Linear- Column Hinge Model 3 Linear

Figure 9. Effects of Column Removal on Pushdown Analyses of Prestressed Girders

13
221 Model 5 with Damaged Web

222 Another damage scenario considered is the failure in the prestressing tendons in one of the webs.
223 The case is represented by setting the flexural stiffnesses theoretically as zero along the length of
224 the bridge. The second girder in the grillage model is selected as the failed girder to simulate a
225 worst case scenario because the maximum demand in the originally intact system is expected on
226 this girder due to the position of the two HS-20 trucks. Due to the interior girder loss, the exterior
227 girder becomes the most critical component when the structure is subjected to the eccentric
228 vehicle load. Figure 10 shows the ultimate capacity of the system or collapse for Model 5 is
229 reached when the total load LFd = 18.2. Thus, the redundancy ratio, Rd, is equal to 18.2/18.0
230 =1.01.

Model 5: Vertical Pushdown with Girder Losses


30

25
Number of Design Trucks

20

15

10

0
0 1 2 3 4
Mid-Span Vertical Displacement (ft)
Model 3: Original Case Model 5: Int. Girder Loss Model 5: Linear

Figure 10. Pushdown Analysis of the Bridge with Web Damage

231 As reported in NCHRP Project 12-86 Interim Report 1 and NCHRP Report 406 Appendix D,
232 grillage modeling methods in this study have been verified by experimental tests of four concrete
233 box-girder bridge models that were tested by Kurian and Menon (2007), a prestressed concrete
234 box, tested by Mirza et al. at McGill University (1990) and a three-column bridge bent tested by
235 McLean et al (1998). The results demonstrate that the grillage model can be used to pred ict the
236 collapse loads of concrete box girders and column bents with a good accuracy.

14
237 Redundancy Ratios and System Factors
238
239 Table 1 summarizes the redundancy ratios for the different models. According to NCHRP 406, a
240 redundancy ratio for the originally intact bridge subjected to overloading should produce a
241 redudancy ratio LFu/LF1 greater than 1.30 to be considered suffciently redundant. Damaged
242 bridges should give LFd /LF1 ratio of 0.50 or higher.

243 The redundancy ratios compare the maximum capacity of the sys tem to that of the first member
244 to fail. As an example, according to linear elastic pushover analysis of Model 1, the lateral force
245 that causes the first member to reach its load carrying capacity P1 is 501kips when the bottom
246 section of the pier reaches its flexure capacity. Following a nonlinear pushover analysis, the
247 ultimate capacity of the system or collapse is reached when the total load P u=720 kips.
248 According to linear “pushdown” analysis of Model 3, if two HS-20 trucks are loaded in the
249 middle span, the first member fails in positive bending when the weight of these trucks is
250 incremented by a factor LF 1 =18.0. When the nonlinear pushdown analysis is executed, collapse
251 takes place when HS-20 load is multiplied by a factor LFu=24.1. The results of Model 1 through
252 3 in Table 1 show that this bridge provides good levels of redundancy for the ultimate limit state
253 due to either vertical or lateral overloading of the originally intact bridge. Although lateral
254 loading of systems with pinned connections between superstructures and substructures shows 8%
255 lower redundancy ratio compared to systems with integral connections, the bridge system with
256 pinned connections still provides adequate level of redundancy. Further studies show that in the
257 longitudinal loading analysis, the integral connection allowed for similar redundancy levels.
258 However, for the cases where the columns are connected to the superstructure through pinned
259 supports, the redundancy level is vastly reduced. The results of Model 4-a, 4-b and Model 5 also
260 show significant higher redundancy ratios than the minimum redundancy value of 0.50 for
261 damaged scenarios under vertical load. This multi-cell bridge satisfies all the redundancy criteria
262 provided in NCHRP 406. However, single cell box girder bridges are non-redundant due to
263 equally loaded girder webs.
264
265 A reliability analysis is used to calibrate the system factors, based on the results of the pushover
266 and pushdown analyses. The reliability calibration details are described in Chapters 2-5 of
267 NCHRP Report 776. Based on the redundancy ratios in Table 1, recommended values for the
268 system factors  s for typical single cell and multi-cell prestressed concrete box girder bridges
269 under lateral loads are given in Table 2. In Table 2, the Risk factor,  , reflects the uncertainties
270 in estimating the load demand and capacity; 1.10 is a multi-column factor for two-column
271 systems; 0.24 is a curvature factor;  u is the ultimate curvature of the weakest column in the
272 bent which can be obtained from a cross section analysis;   is the curvature correction factor for
273 cases with weak connecting elements and weak details (see Equation 3); 3.64 10 4 (1/ in.) is the
274 average curvature for a typical unconfined column and 1.55 103 (1/ in.) is the average
275 curvature for a typical confined column. One column bents and a two-column system with one
276 column loss are considered as a non-redundant system. For non-redundant systems, system
277 factors of 0.75 and 0.85 are proposed for seismic hazards and all other lateral loads, respectively.
278 Details can been found in Chapter 4 of NCHRP Report 776.

15
279 Table 3 lists the recommended system factors  su and  sd for typical single cell and multi-cell
280 prestressed concrete box girder bridges under vertical loads. In Table 3, a system factor of 0.80 is
281 applied to the resistance side of the AASHTO LRFD Bridge Design checking equation (Equation
282 1) to penalize the single cell box girder bridges by requiring their members to provide higher
283 safety levels than those of similar bridges with redundant configurations. While for the multi-cell
284 box girder bridges, a system factor  s=1.0 is used for originally intact bridge due to adequate
285 redundancy levels and a higher system factor  s=1.2 is applied for the significantly redundant
286 bridge systems with one column loss or one web damage.

Table 1 Summary table for the redundancy ratios of the prestressed box girder bridge

Analysis Case Model Pu/P1 LFu/LF1 LFd/LF1 Loading Initial Damage

Model 1 1.44 --- --- Lateral None

Model 2 1.32 --- --- Lateral None

Model 3 --- 1.34 --- Vertical None

Model 4-a --- --- 1.32 Vertical One column

Model 4-b --- --- 1.06 Vertical One column

Model 5 --- --- 1.01 Vertical One web

Table 2. System factors for single cell and multi- cell box girder under lateral load

Variable Applicability Recommended value

Seismic hazards  =0.75


 , risk factor
All other lateral loads  =0.85
Non-redundant s  
 s, system factor
systems

  u  3.64 104 (1/ in.) 


s   1.10  0.24  
 s, system factor Redundant systems 
3 4
1.55 10 (1/ in.)  3.64 10 (1/ in.) 

16
287 The correction factor   is given as:

288
M available  M p column
  if M u column  M available  M p column
M u column  M p column
u connection
289   if M available  M u column and u connection  u Eq. (3)
u
   1.0 if M available  M u column and u connection  u
system is non-redundant if M available  M p column

290 where M available = moment capacity of the connecting elements such as cap beams and
291 pile caps or the reduced moment that can be supported by the column based on the available
292 shear reinforcement, development length, splice or connection detailing. Details on how to
293 calculate the available moment capacity for a member with weak detailing are available in the
294 FHWA Seismic Retrofitting Manual for Highway Structures, Part 1 Bridges,

295 M p column = plastic moment capacity of column,

296 M u column =ultimate overstrength moment capacity of column calculated using


297 nonlinear sectional capacity analysis programs or conservatively estimated to be 1.15 M p column ,

298 u = ultimate curvature of the weakest column in the bent,

299 u connection = minimum ultimate curvature of the connecting elements.

Table 3. System factors for single cell and multi- cell box girder under vertical load

Intact system
Bridge cross section type System factor

Single cell box girder bridges su  0.80

Multi-cell box girder bridges su  1.00

System in damaged state condition


Bridge cross section type System factor

Single cell box girder bridges sd  0.80

Multi-cell box girder bridges sd  1.20

17
300 Conclusions
301
302 This paper investigated the redundancy of single cell and multi-cell prestressed concrete box
303 girder bridges subjected to vertical vehicle loads or lateral loads. For the prestressed concrete
304 box girder bridge systems under vertical loads, the analyses demonstrate that multi-cell box
305 girder bridges can show adequate levels of redundancy in their originally intact conditions and
306 significantly redundant bridge systems with one column loss or one web damage according to the
307 criteria proposed in NCHRP Reports 406 and 776 assuming that the bridge members have been
308 designed to satisfy the applicable specifications. However, single cell box girder bridges are non-
309 redundant due to equally loaded girder webs. For the bridge systems under lateral loads, one
310 column bents and a two-column system with one column loss are considered as non-redundant
311 systems.
312
313 This paper proposed a set of system factors for prestressed concrete box bridges accounting for
314 the structural redundancy. The proposed system factors can be used during the design and safety
315 assessment of prestressed concrete box girder bridges subjected to transient lateral loads or
316 vertical vehicle loads.

317 Acknowledgements
318
319 Support for this study by National Cooperative Highway Research Program (NCHRP) through
320 Project NCHRP 12-86 is gratefully acknowledged. The writers are also grateful for the support
321 and guidance provided by Dr. Waseem Dekelbab, senior program officer and the NCHRP 12-86
322 project panel. Special thanks go to Mr. David Beal and Mr. Bala Sivakumar from HNTB for
323 their reviews, advice and valuable comments.

18
324 References
325
326 AASHTO (2012). Load & Resistance Factor Design (LRFD) Bridge Design Specifications. 6th
327 edition, American Association of State and Highway Transportation Officials, Washington,
328 DC.
329 AASHTO MBE-2-M (2011). Manual for Bridge Evaluation, 2nd Edition, American Association
330 of State and Highway Transportation Officials, Washington, DC.
331 ASCE-AASHTO Task Committee, (1985). “State of the Art Report on Redundant Bridge
332 Systems”, Journal of Structural Engineering, ASCE, Vol. 111, No. 12.
333 Buckle, I., Friedland, I., Mander, J., Martin, G., Nutt, R., Power, M. (2006). Seismic Retrofitting
334 Manual for Highway Structures, Part 1 Bridges, FHWA-HRT-06-032, Turner-Fairbanks
335 Highway Research Center, McLean, VA
336 CAN/CSA-S6-06, Canadian Highway Bridge Design Code (2006). Canadian Standards
337 Association, Canada.
338 Dexter, R J, Connor, R J, Mahmoud, H (2005). Inspection and Management of Bridges with
339 Fracture Critical details, NCHRP Synthesis Report 354, Transportation Research Board,
340 Washington DC.
341 Florida DOT Bridge Load Rating Manual (2012). Topic No. 850‐010‐035, Florida Department
342 of Transportation.
343 Frangopol, D.M., and Nakib, R. (1991). “Redundancy in highway bridges”, Engineering Journal,
344 American Institute of Steel Construction, 28(1), 45-50.
345 Ghosn, M. (2010). Bridge System Safety and Redundancy, National Cooperative Highway
346 Research Program, NCHRP Interim Report 1, Transportation Research Board, National
347 Academy Press, Washington DC.
348 Ghosn, M., and Moses, F., (1998). Redundancy in Highway Bridge Superstructures. National
349 Cooperative Highway Research Program, NCHRP Report 406, Transportation Research
350 Board, National Academy Press, Washington DC.
351 Ghosn, M. and Yang, J., (2014). Bridge System Safety and Redundancy, National Cooperative
352 Highway Research Program, NCHRP Report 776, Transportation Research Board, National
353 Academy Press, Washington DC.
354 Hambly E.C. (1991). Bridge Deck Behavior. London: Chapman and Hall, Ltd.
355 Hovell CG. (2007). Evaluation of Redundancy in Trapezoidal Box Girder Bridges Using Finite
356 Element Analysis. (Master thesis, University of Texas at Austin).
357 Hunley, C.T., and Harik, I.E. (2012). “Structural Redundancy Evaluation of Steel Tub Girder
358 Bridges.” Journal of Bridge Engineering, 17(3), 481-489.
359 Kurian B, Menon D. (2007). Estimation of Collaspe Load of Single-cell Concrete Box-Girder
360 Bridges. Journal of Bridge Engineering, vol. 12, pp. 518-526.
361 Liu, D., Ghosn, M., Moses, F., and Neuenhoffer, A., (2001). Redundancy in Highway Bridge
362 Substructures. National Cooperative Highway Research Program, NCHRP Report 458,
363 Transportation Research Board, National Academy Press, Washington, DC.
364 McLean, D.I., Kuebler, S.E., Mealy, T.E., (1998) Seismic performance and retrofit of
365 multicolumn bridge bents, Washington State University. WA.
366 Mertz D. (2006). LRFR: FDOT Rating Policies & Procedures, Presentation at Summer 2006
367 FDOT, Design Conference. Orlando, Florida.
368 Mirza, M. S., et al. (1990). An experimental study of static and dynamic responses of prestressed
369 concrete box girder bridges. Canadian Journal of Civil Engineering, 17(3), 481-493.

19
370 Yang, J., & Ghosn, M. (2015). System Factors for Evaluating the Redundancy of Bridge Systems
371 under Vertical Load. In Transportation Research Board 94th Annual Meeting, No. 15-1039,
372 Transportation Research Board, National Academy Press, Washington, DC.
373 Yang, J., & Ghosn, M. (2015). System Factors for Evaluating the Redundancy of Bridge Systems
374 under Lateral Load. In Transportation Research Board 94th Annual Meeting, No. 15-3516,
375 Transportation Research Board, National Academy Press, Washington, DC.
376 Zokaie T, Osterkamp TA, Imbsen RA. (1991) Distribution of Wheel loads on Highway Bridges.
377 Final Report, NCHRP project 12-26; Transportation Research Board, The National
378 Academies, Washington, DC.

20

You might also like