strategies for load rating of infrastructure populations

47
1 Strategies for Load Rating of Infrastructure Populations: A Case Study on T-beam Bridges F. N. Catbas 1 , S. K. Ciloglu 2 , A. E. Aktan 3 ABSTRACT: There is consensus on the importance of objectively and reliably assessing the condition and load capacity of aged bridges. Although each bridge may be considered as a unique structure, the behavior of many bridge types may be governed by only a few mechanisms and related parameters, especially if a population is constructed from standard designs. By identifying these parameters, and their variation within the population, it is possible to extend findings such as load rating obtained from a statistical sample to the entire population. Bridge type-specific strategies for load rating and condition assessment in conjunction with statistical sampling may therefore offer significant advantages for inspecting and load rating bridges sharing common materials, similar geometry and detailing, and the same critical behavior mechanisms. In this paper, the writers present their recent work on load rating of the reinforced concrete T-beam bridge population in Pennsylvania to objectively re-qualify them based on field-calibrated finite element models. Keywords: Load rating, distribution, bridge, population, field test, finite element model 1 Assistant Professor, Civil and Environmental Engineering Department, University of Central Florida P.O. Box 162450, Orlando, FL 32816-2450, Phone: 407-823-3743; e-mail: [email protected] (corresponding author) 2 Ph.D. Student, Drexel Intelligent Infrastructure and Transportation Safety Institute, Drexel University, 3001 Market Street, Suite 50, Philadelphia, PA 19104 3 John Roebling Professor of Civil Infrastructure Studies, Director of Drexel Intelligent Infrastructure and Transportation Safety Institute, Drexel University, 3001 Market Street, Suite 50, Philadelphia, PA 19104

Upload: others

Post on 10-Jan-2022

0 views

Category:

Documents


0 download

TRANSCRIPT

1

Strategies for Load Rating of Infrastructure Populations: A Case Study on T-beam Bridges

F. N. Catbas1, S. K. Ciloglu2, A. E. Aktan3

ABSTRACT: There is consensus on the importance of objectively and reliably assessing the condition

and load capacity of aged bridges. Although each bridge may be considered as a unique structure, the

behavior of many bridge types may be governed by only a few mechanisms and related parameters,

especially if a population is constructed from standard designs. By identifying these parameters, and their

variation within the population, it is possible to extend findings such as load rating obtained from a

statistical sample to the entire population. Bridge type-specific strategies for load rating and condition

assessment in conjunction with statistical sampling may therefore offer significant advantages for

inspecting and load rating bridges sharing common materials, similar geometry and detailing, and the same

critical behavior mechanisms. In this paper, the writers present their recent work on load rating of the

reinforced concrete T-beam bridge population in Pennsylvania to objectively re-qualify them based on

field-calibrated finite element models.

Keywords: Load rating, distribution, bridge, population, field test, finite element model 1 Assistant Professor, Civil and Environmental Engineering Department, University of Central Florida

P.O. Box 162450, Orlando, FL 32816-2450, Phone: 407-823-3743; e-mail: [email protected] (corresponding author)

2 Ph.D. Student, Drexel Intelligent Infrastructure and Transportation Safety Institute, Drexel University, 3001 Market Street, Suite 50,

Philadelphia, PA 19104

3John Roebling Professor of Civil Infrastructure Studies, Director of Drexel Intelligent Infrastructure and Transportation Safety Institute,

Drexel University, 3001 Market Street, Suite 50, Philadelphia, PA 19104

2

1 INTRODUCTION

1.1 Statement of the Problem

It is generally acknowledged that most critical infrastructure systems have been falling short of providing

satisfactory operational performance under everyday demands, and their constructed elements have been

appraised to have poor structural conditions (ASCE 2003, Report Card for America’s Infrastructure).

Highway bridges are critical nodes of the highway transportation network. Nearly 30% of the entire U.S.

bridge population (592,246 bridges) in the National Bridge Inventory (NBI) has been reported as

“structurally deficient or obsolete” based on their “condition rating” and factors such as “posting”

regardless of size or importance (NBI 2003). The bridge population continues to age, and we lack the

funds for immediate rehabilitation or renewal of existing bridges that are deemed as “structurally deficient

or obsolete.” As we defer rehabilitating or replacing posted bridges due to financial constraints, it has

become even more important to be able to objectively evaluate the structural condition and safe load

capacity of bridges and prioritize their replacement in an integrated asset management framework. In the

last decade there has been a great thrust for objective condition assessment, repair and renewal

technologies, and non-destructive evaluation methods. However, it is not a realistic expectation to have

the time and resources for an in-depth evaluation of every single one of over 150,000 bridges deemed

“structurally deficient or functionally obsolete” (Chase 2001).

1.2 Review of Current Practice

The National Bridge Inventory (NBI) contains 116 data fields for each bridge irrespective of the bridge

type, importance and other possible distinctions. There are three data fields containing information for the

structural condition rating, five data fields for the appraisal ratings, and several fields for the general

attributes of a bridge. The structural condition is mainly evaluated through data from the biennial

inspections that are conducted in accordance with the guidelines set forth in the National Bridge

Inspection Standards (NBIS) (NBI 1998, 2003, FHWA 1995).

3

Only a limited amount of structural condition information exists in the NBI, consisting of condition

ratings based on visual inspection. The NBI contains three data fields for condition rating which are

described in the Recording and Coding Guide for the Structure Inventory and Appraisal of the Nation’s

Bridges (FHWA 1995). The coding descriptions are very general and subjective. A recent study by Federal

Highway Administration (FHWA, 2001) has indicated that on the average at least 78% of the average

Condition Ratings from visual inspections were incorrect with a 95% probability if FHWA Non-

destructive Validation Center (NDEVC) reference condition ratings are assumed correct (FHWA-

NDEVC, 2004). The variations are a result of factors such as the inspectors’ experience, type of bridge

and condition of the bridge. It should be noted that some states such as Pennsylvania have instituted

additional quality standards for consistency in bridge inspections. In addition, some states, such as

California, have incorporated a more detailed element-level inspection and recording program for their

bridge management (Roberts and Shepard 2001).

In many bridge management programs, condition rating is complemented with load capacity rating, where

the latter is typically obtained as described in the American Association of State Highway and

Transportation Officials (AASHTO) “Manuals” for Condition Evaluation and for Strength Evaluation

(AASHTO 1989, 2000). Most transportation departments utilize software such as BAR7 (PennDOT,

2001) to compute capacity and demand as defined by the “Manuals.” Typically, highly idealized analysis

methods are used, resulting in conservative load rating results. The “Manuals” support various indirect

manners for incorporating inspection results into load rating such as by modifying the impact factor for

demand based on the wearing surface condition evaluation, or by modifying the resistance factors based

on field inspection or maintenance activities. However, even the best practice cannot objectively link “as-

is condition” and “load capacity rating” of a bridge. Such a linkage, however, is possible through

“structural identification” (Aktan et al, 1997, 1998).

4

1.3 Evaluation of Large Populations

To restructure the problem of bridge condition assessment and effective maintenance management by

taking advantage of objective experimental data from the field, the concept of statistical sampling offers

great promise. For example, airplane fleet owners take advantage of common symptoms and in-depth

inspections of just a few members of a fleet and effectively extrapolate these to large populations of

similar vehicles sharing a standard design, use-history and age. Many bridge engineers and managers have

viewed every highway bridge as a distinct and unique structure. In spite of considerable variation in

material properties, geometry, structural details and visual appearance, the load resisting

mechanisms and critical failure modes of most bridge populations may be governed by only a limited

number of independent parameters. It should be possible to classify bridge populations by establishing

the critical parameters for load capacity and failure mode by properly designed and executed research. This

would permit the evaluation and management of a large bridge population by selecting and studying a

statistical sample. In this manner, although every individual bridge may still be inspected as a distinct

structure for bridge-specific critical problems, bridge managers may take advantage of information from

statistical samples to manage bridges more effectively.

It should also be noted that the concept of statistical sampling of large bridge populations has been

implemented by other researchers. Livingston and Amde (2000) investigated the causes of micro-cracking

and additional deterioration in concrete due to formation of mineral ettringite by analyzing bridge

populations. Madanat et al. (1997) developed statistical models for infrastructure facility deterioration by

including the presence of persistent facility-specific but unobserved factors such as construction quality.

They then extended the model to investigate the presence of state dependence to develop a model for

bridge-deck deterioration. The data used for their study consisted of 5,700 state-owned bridges in Indiana

and the condition ratings of these bridges were included in the analysis.

5

1.4 Re-Qualification of T-beam Bridges in Pennsylvania

The total number of single span T-beam bridges in Pennsylvania is 1,899 and approximately 60% of this

population is older than 60 years, with a maximum age of 101 years. Approximately 90 of these bridges

have been posted. Bridge engineers have intuitively sensed that even after aging and deterioration, cast-in-

place RC T-beam bridges with sound abutments inherently possess a greater load capacity than what their

current BAR7 rating give and their low condition ratings may imply. However, there has not been a

scientifically proven method to confidently evaluate the impacts of accumulated deterioration and damage

on the safe load capacity of a bridge.

1.5 Objectives and Scope

The writers have been exploring how to develop and implement condition assessment and load rating

strategies for recurring types of structures by taking advantage of the statistical sampling concept, which

makes it possible to have a better understanding of the bridge performance and to develop more effective

and practical methods to manage a specific bridge-type. They also anticipate that this concept can be

implemented on different bridge types as well as other infrastructure systems. The main objective of this

paper is to present the statistical sampling approach applied to large populations of recurring bridges and

the use of experimental, analytical and information technologies on the statistical sample for objective

condition assessment of the entire population. An overview of the issues related to this approach is

provided along with the various experimental and analytical technologies. These technologies are

employed as a complement to visual inspection results, providing objectivity to the current practice for

better operation and maintenance management of large populations of structures with similar geometric

and condition parameters. To illustrate the application and implementation on real life structures, the

findings and results from a recent research study to re-qualify the load rating of single span reinforced

concrete T-beam bridges in Pennsylvania are presented. In this context, the scope of the paper is given as

follows: 1) Discuss and provide an example on the statistical sampling strategy for the management of a

T-beam bridge population for objective, quantitative bridge load condition evaluation and capacity

6

assessment; 2) Discuss possible methods to determine the contribution of critical structural mechanisms

affecting load condition by means of objective measurements, analytical modeling and parameter

sensitivity studies that can be implemented on statistical populations; 3) Discuss how current visual

inspections and load rating procedures can be complemented and improved using experimental, analytical

technologies and statistical sampling strategies.

2 RESEARCH APPROACH

Research approach included:

1) Statistical evaluation of the entire T-beam bridge population as a fleet analogous to a truck or

an aircraft fleet: By identifying a representative sample that reliably represents the critical

relevant characteristics of the entire population (Ang and Tang, 1975), and by investing in

instrumentation, testing and monitoring of the statistical sample, reliable management decisions

may be reached for the entire fleet. The use of statistical sampling has been common for polling;

the use of this approach has been debated extensively for census, and in fact it was proven as a

more reliable approach than attempting a one-by-one headcount (NY Times, 2002).

2) Observations and experiments on the bridges in the statistical sample, and analytical studies

with field calibrated 3D FE models: These studies helped determine the most critical condition

and nominal structural parameters and helped establish the actual condition of the T-beam bridge

population in terms of objective parameters such as strain and deflection influence coefficients

and their measured values.

3) The socio-technical factors governing the determination and use of the load capacity rating in

bridge management: The highest acceptable load capacity rating that would conform to the

inherent conservatism in the current AASHTO specifications and the uncertainty in the visual

inspections is identified. Effective load distribution factors for use with simple-beam models for

analysis were formulated for Pennsylvania Department of Transportation (PennDOT) engineers to

7

be able to compute the highest utilizable load rating of any T-beam bridge without violating

AASHTO specifications. The effective load distribution formulations developed for the RC T-

beam bridges should be of considerable value for all state departments of transportation

following review by AASHTO. However, researchers see a need and anticipate that AASHTO will

agree to further verification by destructive tests of several decommissioned T-beam bridges.

Properly designed and executed destructive tests, accompanied by appropriate nonlinear finite

element (FE) analyses would be needed to confirm that even the extreme levels of deterioration

and any loss in the secondary elements or boundary restraint mechanisms would not affect the

minimum expected level of serviceability, safety and reliability from RC T-beam bridges and

more importantly, would not lead to undesirable failure modes.

3 STATISTICAL SAMPLING OF BRIDGE POPULATIONS

A main objective of this study is to present that structural identification of a statistically representative

sample of a bridge population may be used for objectively and reliably characterizing the entire

population. For example, an authority may classify its ten thousand steel-stringer bridges into various

groups of several thousand each, depending on the statistically independent parameters that govern the

load capacity rating and other concerns that are taken into consideration for bridge management. The

bridges making up a sample may be rigorously inspected and tested by expert bridge engineers, creating a

sufficient amount of data and insight for the management of the entire group for decades to come. In this

manner, it is possible to take maximum advantage of the bridge-type specific heuristics that has been

accumulated, and to integrate this with the advanced technological tools that offer reliable and

measurement-based determination of serviceability and load capacity.

3.1 Statistical sampling of T-beam bridges in Pennsylvania

The writers selected a statistical sample of single span, RC T-beam bridges in Pennsylvania. Although a

large number of the T-beam bridges are aged and deteriorated, anticipated to be nearing the end of their

8

service life, it is also realized that the actual load capacities and structural condition of these bridges may

be much greater than the estimates due to the desirable inherent qualities of cast-in-place RC beam-slab

systems. The entire reinforced concrete single span T-beam bridge population in PA consists of

approximately 1,899 RC T-beam bridges and these bridges were constructed mostly between 1900’s and

1960’s. Most of the bridges were constructed using standard set of drawings (Standards, 1983). As most

of these T-beam bridges share geometry and design details, materials and similar cast-in-place

construction, and since recent field experiments on monolithic cast-in-place RC beam-slab behavior

demonstrated excellent reserve capacity (Al-Mahaidi et al. 2000 and Song et al. 2002), this bridge

population was an excellent candidate for implementing statistical evaluation approaches. A statistical

study was conducted on 1,651 bridges with complete information in the NBI out of the entire population

of 1,899 single span RC T-beam bridges in Pennsylvania. We note that the entire population of T-Beam

bridges including multi-span bridges is 2,384 in PA and 37,408 in the USA (NBI, 2003). The findings

from the study are expected to be useful for making decisions on the entire population.

In this study, the writers assume that the load capacity of the RC T-beam bridge population can be

considered to be a function of “nominal structural,” and “as-is condition” parameters.

The governing independent nominal structural parameters are established as the span length and the

skew angle. T-beam bridges were constructed using a standard set of drawings, the majority dating back to

the 1930’s (Standards, 1983). In the standard design drawings, the element dimensions and reinforcement

details are dependent only on the span length and skew of the bridges. For example, when a bridge with

certain plan geometry is selected, the beam sizes, reinforcement and all other details such as parapets,

diaphragms, support details, etc. are automatically established. This “mechanistic” dependency greatly

reduces the number of independent structural parameters.

The governing independent condition parameters are the location of a bridge, its age, its current

condition rating and input from District Engineers regarding the most deteriorated bridges that they are

9

concerned with. The challenge was in identifying which of the many possible parameters were dependent

on others, and those that impacted the actual load carrying capacity of a bridge. Different parameters were

analyzed by manipulating the inventory records by the help of GIS software ArcVIEW (2001). Statistics,

histograms, population characteristics and geographic distribution of the bridges within the state of

Pennsylvania were evaluated (Catbas et al, 2002, 2003). The parameters that are used for the statistical

identification incorporate population density, density of bridges within a geographic region, the

geographic/climate distribution and any related socio-technical factors such as the personnel resources of

the District with jurisdiction over a geographic region. Geographic distribution of the selected statistically

representative sample of 60 bridges along with the entire population is shown in Figure 1. Additional

parameters and their distribution with respect to the entire population are presented in Figure 2.

In the course of one year, the writers visited 27 bridges within the T-beam bridge population sample for

in-depth visual inspection, material sampling, non-destructive evaluation (NDE) studies, and to confirm

that the independent parameters that governed load capacity were indeed those that were stipulated in the

sampling process. Of the sample, four bridges were then subjected to in-depth structural testing and field-

calibrated finite element modeling for structural identification so that reliable simulations for sensitivity

studies of load capacity could follow. The results validated the potential of the fleet concept, and were

sufficient for reaching several recommendations that promise significant impact on bridge management.

The entire study was completed in two years.

4 EXPERIMENTAL AND ANALYTICAL STUDIES ON THE BRIDGE POPULATION

4.1 Preliminary Analytical Studies and Field Evaluations

4.1.1 Finite Element Modeling of a Typical Bridge

A bridge representing an average geometry and condition is selected for FE solid modeling and analysis

based on a fine mesh. Analysis results helped to determine the contribution of different structural

elements and mechanisms to load rating. Different levels of deterioration and damage are simulated to

10

investigate their impact on load rating. These findings from preliminary analyses of an average bridge are

used to design and conduct field inspections to document the as-is conditions of 27 bridges.

4.1.2 Field Evaluations of Bridges

In-depth field investigations including detailed deterioration and damage mapping, imaging, coring, and

condition documentation with a focus on the critical areas, elements and mechanisms such as the

boundary restraints that affect load rating. Inspection results made available by Pennsylvania Department

of Transportation are used as a guide during the field evaluations. A data-base is constructed for managing

the information corresponding to approximately one-half of the initial sample of 60 bridges. The reduced

sample-size of 27 bridges also included some bridges with the worst condition ratings.

4.2 In-depth Testing of Four Sample Bridges

In-depth field tests included extensive instrumentation and truck load tests, impact modal testing using an

impact-hammer, and Falling Weight Deflectometer (FWD) testing. The dynamic tests were conducted to

verify the global mass and stiffness distribution characteristics. In addition, using FWD and modal tests on

T-beam bridges to extract flexibility coefficients as condition indices were explored on T-beam bridges.

Four T-beam bridges were studied in this manner for objective data to quantify the actual operating

stresses and behavior of the bridges in their as-is conditions. The test results are processed to determine

the bridge dynamic response frequencies and mode shapes, critical concrete and steel strains and

maximum deflections under various levels and configurations of live load. The four bridges covered the

spectrum of geometric and condition parameters.

4.3 Finite Element (FE) Model Calibration by Field Test Data

4.3.1 Construction and Calibration of FE Models

3D FE models are calibrated to simulate the actual geometry and as-is material, continuity and boundary

conditions of a structure. Such models may provide much more reliable estimates of actual load capacity

11

rating of a bridge than idealized simple beam models. However, the process of first constructing and then

calibrating a 3D FE model using both dynamic modal analysis and static load test measurements requires

considerable expertise and may not be routinely implemented in practice although feasible in the context

of evaluating a sub-set of the statistical sample.

4.3.2 Field calibrated modeling of four test bridges

Based on the statistical study (Catbas et al, 2003), four bridges in the sample population were selected for

detailed investigations, including 3D FE modeling, field testing and FE model calibration by field test

data. The FE models of the bridges are initially developed using the nominal structural and condition

parameters, and these are then calibrated based on the field inspection, NDE, material test results and

structural load test results. The field-calibrated models are then analyzed for load rating by simulating two

side-by-side HS20-44 trucks positioned for maximum moment and shear, respectively. The results are

compared with those obtained by using BAR7 structural analysis software.

5 RESULTS FROM FE MODEL STUDIES AND FIELD EVALUATIONS

Writers tested and analyzed several bridges. One of the test bridges, the Swan Road Bridge, is selected to

serve as an example, and the results from other bridge tests are also reported later in this paper. One of the

main reasons of selecting Swan Road Bridge is that the geometric and condition characteristics of the

Swan Road Bridge represent an almost population average. Swan Road Bridge was constructed in 1937,

is 26 ft (7.9 m) long, has no skew and is 26 ft (7.9 m) wide supported on 6 T-beams. The condition rating

of its superstructure is 6 (Figure 3). A typical 3D FE model that is constructed using solid elements and

frame elements available in the library of the SAP 2000 V8 software (2002) is illustrated in Figure 4.

Each reinforcing bar and its bond with concrete are explicitly simulated. Such a fine microscopic approach

to 3D geometric-replica analytical modeling is now practical and enables explicitly simulating every

material point of the bridge for an accurate representation of the geometry, the actual behavior

mechanisms and a wide range of possibly existing deterioration or damage.

12

An example snapshot illustrating the load testing of the Swan Road Bridge is shown in Figure 5 followed

by the relevant experimental results. Figure 5 includes the instrumentation plan for the static and crawl

speed load tests. Linear Variable Displacement Transducer (LVDT) sensors are used to measure

displacements. Weldable strain gages which are microdot welded to reinforcement are used for rebar

strain measurements and clip gages are used for concrete strain measurements. These sensors are installed

under the bridge and the respective locations of the sensors are shown in the figures. In addition, 12-15

accelerometers are mounted on the deck to measure the dynamic properties of the bridges.

The test results from the impact test, Falling Weight Deflectometer and the load testing are summarized in

Tables 1 and 2. For global calibration of the FE models, the results of the dynamic tests were compared

with the finite element eigenvalue analysis results. The frequencies of the nominal models for both

bridges are lower than the measured frequencies, indicating that the analytical models simulate a greater

flexibility than actual. After calibration, especially of the existing boundary conditions, the errors in

frequencies for the first three global modes of the models were reduced by more than 50% (Table 3). It

should be noted that although a "100% match" between all experimental data and analytical models for

real life structures cannot be expected, an even better correlation than summarized in Table 3 can be

achieved by conducting a parameter study at a microscopic level, such as by modifying the input

properties for various finite elements and each abutment seating separately. However, writers experience

has been that such a microscopic level fine-tuning of the models for correlating strains would not have

any significant impact on the calculated load rating factors. The correlations between the measured and

simulated deflections and strains under static loads before and after calibrating the 3D FE models are

presented in Figure 6.

The most significant modification that was required during the calibration of the FE models was

incorporating the actual restraints at the boundaries, between the stiff diaphragm beams sitting on and

connected by dowels and expansion plates anchored to the sub-structure. The lateral compressive thrust

13

exerted on the bridge by the pavement proved to be an additional mechanism at the boundaries (Figure 4)

significantly affecting load rating. When the 27 bridges were visually inspected, the condition of

continuous pavement that provides lateral thrust was observed and documented.

In order to determine the contribution of various mechanisms and to evaluate the impact of extreme

damage (Figure 7), parametric sensitivity analyses were conducted using the 3D FE model. The results are

summarized schematically in Figure 8. The analytical study summarized in the following serves to:

1) Compare the load capacity ratings based on an idealized modeling of the bridges by a simple beam

free-body and analyzed by the BAR7 software with those determined by analyses of field

calibrated 3D FE models.

2) Estimate possible changes in load ratings in the event of possible extremes of unmitigated

deterioration and damage that may occur during, say, the next five to ten years.

3) Evaluate the 3D FE analysis results for deriving conclusions regarding the possible impacts of

critical material properties, structural elements and load distribution mechanisms on the load

capacity ratings of these two bridges. The conclusions derived for the four bridges (one is given as

an example in the paper) are then qualitatively generalized to the broader population.

5.1 Flexural Load Rating Results

1) Figure 9 indicates that the BAR7 analysis of the Swan Road Bridge yields rating factors of 1.27 for

flexure, and this is indeed the current load ratings for these bridges in PennDOT District 6 records.

2) The load ratings for flexure based on the field calibrated 3D FE model is 3.18 (150% higher than the

BAR7 load rating) for the Swan Road Bridge. It is important to note that the calibrated FE models

incorporate a reduced elasticity modulus for concrete and simulate all of the deterioration that was

identified during field inspections. The corresponding load rating values are still much higher than the

load rating results based on BAR7, although the latter do not incorporate any deterioration.

14

3) In the event of possible extreme deterioration, simulated by reducing the concrete and steel of the

beams in the field-calibrated models, the load rating is reduced to 2.11 for the Swan Road Bridge.

This value is still 66% higher than its counterpart based on BAR7 analysis. Simulated deterioration

does not affect the demands significantly, and in fact somewhat attenuates the maximum demands

while the reductions around 30% are mainly caused by the reductions in the capacity due to loss of

material.

4) The results presented in Figure 9 indicate that the boundary conditions have the most significant

impact on the load ratings. The field tests revealed that friction and dowels between the stiff lateral

diaphragm beams of the superstructure and the beams on the abutments create a very effective

restraint, prohibiting any slippage and other movements. Lateral soil pressure and pavement thrust

further slightly contribute to the restraint. When these effects are ignored and the boundary conditions

are changed to pin-roller, the resulting load ratings become 1.99 (compare to 3.05 with pin-pin

boundary conditions) for the Swan Road Bridge. This load rating, however, is still 57% higher than

the BAR7 rating mainly due to the lateral load distribution due to the slab and contributions of the

secondary elements (diaphragms and parapets).

5) The diaphragm beams provide effective rotational restraints (and thereby increased bending stiffness)

at the boundaries, which in turn reduce the critical flexural demand at the mid-span. Similarly,

parapets help distribute the flexural stresses from the mid-span towards the edges by creating very stiff

girders at the edges. When diaphragms are excluded from the model, the load ratings for the bridge

slightly decrease. However, the load rating is still 48% higher than the BAR7 load rating for the Swan

Road Bridge. Using the 3D FE models for calculating the demand for this bridge after neglecting the

restraints at the boundaries, and the contributions of the diaphragms and parapets results in rating

factors of 1.44 for the Swan Road Bridge. This is still 13% higher than their BAR7 rating

counterparts. These increases in load rating are due to fully simulating the distribution of stresses in

the transverse direction due to the slab in the 3D FE model. Figure 9 also reveals that when the

15

probable future deterioration extremes are simulated for the bridges without including any of the

secondary elements and mechanisms, the load ratings for flexure may fall below one (0.88 for Swan

Road Bridge). However, the reduction in load demand due to redistribution would be expected to

yield a load rating higher than 0.88 if a non-linear model is used for concrete.

5.2 Shear Load Rating Results

1) BAR7 analyses yield rating factors of 1.80 for shear for the Swan Road Bridge.

2) 3D FE models effectively simulate the more effective shear distribution due to the presence of the

deck, an effect that is ignored in shear rating by the BAR7 model. The critical shear demands occur

near the supports of T-beams under load at the obtuse angle side of skew bridges. The shear capacity

mechanisms considered in rating included the effective beam concrete, stirrups and bent rebar

contributions similar to the DOT practice. The shear rating for the Swan Road Bridge obtained from

3D FE analysis is 2.69, which is 50% higher than the corresponding rating by BAR7 analysis.

3) Shear rating factors obtained by 3D FE analysis decrease to 2.30 for the Swan Road Bridge, when

extreme probable deterioration is simulated. The dead load and live load demands are obtained from

the FE model in which deterioration is simulated. In addition, shear capacity computation includes the

reductions of concrete, stirrups and bent longitudinal reinforcing bars; as a result, the section capacity

is decreased accordingly. However, shear load rating is still 28% higher than the corresponding shear

rating from BAR7 analysis.

4) Using nominal parameters, pin-pin boundary conditions and including the secondary elements, the

shear rating is 3.54 for the Swan Road Bridge. Because the boundaries are the most critical sections

for shear, an increased stiffness at the boundaries when pin-pin boundary conditions are simulated

result in higher shear demands. When the boundary conditions are changed to pin-roller for the Swan

Road Bridge, the corresponding rating factors for shear become 3.90. The increase in load rating is as

a result of the reduced shear demand at the critical locations when boundary restraints are released.

16

5) Secondary elements in 3D FE models provide an increase in the rating factors by enhancing the

redistribution of stresses and reducing the maximum demand. If both parapets and diaphragms are

ignored and pin-roller boundary conditions are simulated, the shear rating factors become 2.64 for the

Swan Road Bridge. This is still 46% higher than the corresponding shear load rating by BAR7

analysis.

6) When extreme future probable deterioration is simulated in the nominal 3D FE model, the shear

rating factor decreases to 1.97 for the Swan Road Bridge. This is a consequence of the reduced shear

capacity and shear load distribution. The shear rating is still greater than one, and do not govern the

load rating as flexure remains as the critical effect. Shear does not appear to be a concern even when

extreme deterioration is concerned provided that unchecked and hidden deterioration for example due

to alkali-silica reaction is mitigated and the substructures remain in good condition, eliminating

settlements. However, it is important to determine whether shear becomes the prevailing failure mode

due to deterioration, and this is further discussed in the following section.

6 MECHANISMS LEADING TO HIGHER LOAD CAPACITY RATINGS

The analyses presented here for the Swan Road Bridge was repeated for the remaining test bridges which

are shown in Figure 10. The load rating values for the four test bridges obtained by field-calibrated finite

element models and obtained by idealized BAR7 analysis are also compared in the same figure. We

observe that the load rating values based on field-calibrated models range between 3.18 and 5.15, and the

ratio of field-calibrated FE based rating to BAR7 based rating factors vary between 2.50 and 5.10. Based

on these results, it is concluded that the BAR7 approach to load rating of T-beam bridges overestimate the

demand by 2.5 times or larger, and that this conservatism is consistent irrespective of the span, the skew or

condition. During the course of the project it was observed that there is a consistent correlation between

the DOT records of the 27 inspected bridges and their actual condition. The mechanisms that contribute to

higher load rating are summarized in the following.

17

6.1 Demand Mechanisms That Contribute to Higher Load Rating by FEM

The load rating results summarized earlier indicate several mechanisms and parameters that contribute to

a decrease in the load demand, thus considerably enhancing the load rating relative to what is obtained

from a BAR7 analysis. Even after eliminating all of the secondary mechanisms and elements in the 3D FE

model, it is still possible to increase the load rating 13% for the Swan Road Bridge. It is possible to

generalize the mechanisms that reduce critical load demands and lead to rating increases that are not

incorporated in BAR7 analyses:

6.1.1 Importance of Boundary Conditions:

The use of pin-pin boundary conditions may not be justified if this is due to mechanisms such as frozen

bearings. However, when boundary restraints are due to permanent mechanisms such as dowels and lateral

confinement provided by the pavement, and, if the lateral restraints persist during load tests that are

conducted at proof load levels, the use of pin-pin boundary conditions for load rating purposes may be

appropriate. Figure 9(b) and (c) show that the pin-pin boundary conditions provide the largest increase

(53%) in the simulated load capacity rating factors, it is recommended that the boundaries at the super-

and-substructures are carefully inspected and any evidence of movement reported during biannual

inspections.

6.1.2 Lateral Restraint due to Earth Pressure and Pavement Thrust:

The individual T-beams are idealized as simply-supported (pin-roller) when analysis programs such as

BAR7 are utilized. However, as observed during visual inspections as well as identified and simulated in

the field-calibrated FE models, there are effective lateral restraints at the ends of the bridges due to earth

pressure and pavement thrust. The lateral restraints create compressive membrane forces and also increase

the flexural beam stiffness at the boundaries. This effect reduces the maximum span moments.

18

6.1.3 Reinforced Concrete Parapets:

The parapets serve as stiff edge girders along the traffic direction. The corresponding edge stiffness may

have a major effect on flexural and shear distribution. For narrower bridges, the contribution of the stiff

edge girders is more significant than wider bridges. Parapets’ effect on load rating factor may be

quantified based on parameter sensitivity analyses conducted on Swan Road Bridge. Figures 9(d) and (e)

indicate that the load rating factor increased 30% due to the parapets.

6.1.4 Diaphragm Beams:

Lateral and longitudinal movements of actual bridges are restrained due to the dowels and the friction

between the superstructure and substructure at both ends as observed from the field inspections and

indicated by experimental measurements under loads. The lateral diaphragm beams also provide effective

rotational restraints to the superstructure, further reducing the flexural demands at the mid-span. In

addition, the diaphragm beams distribute the reactions along the super-sub structure interface thereby

reducing the shear demand.

6.1.5 Lateral Load Distribution:

In the current load capacity rating practice, an individual beam is taken out as a free-body, idealized as

simply-supported, and the continuity of the bridge in the transverse direction is indirectly accounted for by

means of axle-load distribution factors. This approach is found to significantly underestimate the deck

slab’s contributions to lateral load distribution for many bridge geometries. This contribution is properly

simulated when a properly constructed, geometric replica 3D FE model is used for analysis.

6.1.6 Effective Force Redistribution Due To Cracking:

In general, although the load capacity rating is based on the initiation of yielding in the reinforcement, we

ignore the effects of concrete cracking that occur in advance of yielding. Cracking of concrete is a

mechanism that provides a very effective redistribution of stresses within a T-beam bridge, and therefore

19

effectively reducing the demands and leading to a higher load rating (Shahrooz et al. 1994; Huria et al.

1994).

6.2 Capacity Mechanisms That Contribute to Higher Load Rating

Throughout this study and in deriving the load capacity rating factors in Figure 8, capacity of the T-beam

bridges are computed by strictly following the AASHTO Load Factor Design procedures. The capacity of

each T-beam is calculated as an element separated from the bridge system, and by assuming that flexural

capacity is reached when the first layer of reinforcing steel reaches the nominal yield strain. Shear capacity

is attained when either the stirrups or the bent-bars reach their nominal yield strain. This approach for

computing capacity is well known to underestimate the actual available capacity. Just as many

mechanisms not considered in rating reduce actual internal force demands, there are also many other

mechanisms that are not incorporated in rating but that are known to lead to an increase in capacity. The

actual capacity of T-beam bridges can be better estimated by means of properly conducted non-linear

analysis of 3D FE models, calibrated based on destructive load test data. However, even without

destructive testing or nonlinear FE analysis of an entire bridge, we note the following mechanisms which

are not included in load rating that may considerably increase the actual attainable capacity:

6.2.1 Axial Restraints at the Boundaries:

The axial restraints at the boundaries due to lateral earth pressure and pavement thrust lead to compressive

membrane forces in bridges that induce a multi-axial state of compression in the beam and slab concrete.

Additionally, pin-pin boundary conditions also lead to axial compression in the beams and the slab upon

the deflection of a bridge, termed as the membrane effect. Multi-axial compression due to the lateral

thrust and membrane effect is known to delay the formation of cracking and bond slip, offsetting the

tensile forces in steel, and considerably enhance the compressive strength of concrete relative to what is

obtained from a cylinder test. Therefore, the axial restraints at the boundaries not only reduce demands but

also increase the capacity.

20

6.2.2 Higher Yield Strength and Strain Hardening of Steel:

It is well known that actual reinforcing steel bars have about 25% or greater yield strength than the

nominal strength (ACI 1998). For example, tests on tensile yield strength of Grade 60 rebars with nominal

yield strength of 60 ksi (414 MPa) indicate 133% higher yield strength for about 10% of the test

specimens (MacGregor 1988). Further, steel stress-strain behavior is idealized in load rating as elastic-

plastic. However, at the ultimate, steel stress may be about 40 % higher than at yield due to strain-

hardening. These increases in steel yield stress and maximum strength lead to an increase in the attainable

flexural capacity of an under-reinforced beam by a similar ratio.

6.2.3 Multiple Rebar Layers:

Capacity of the T-beams is computed based on the assumption that the capacity would be attained when

the first layer of rebar reaches the yield strain. When there are additional rebar layers, the yielding of the

rebar layers will be achieved sequentially, and this phenomenon provides redistribution of strains within a

cross-section between different beams. Considerable additional flexural capacity as compared to what is

calculated based on the current assumption that capacity is reached when the rebars at the lowermost layer

yield is attained with multiple rebar layers.

6.2.4 Slab Contribution:

The capacity of beam-slab systems are known to be significantly higher than what is obtained by summing

up the capacity of isolated T-beams. The actual modes of failure observed during laboratory testing of

beam-slab systems have been through a flexural collapse mechanism typically after significant overloads

and excessive deformations are reached following the formation of yield lines (Park and Gamble 2000). In

addition, the shear capacity of T-beam bridges are significantly greater than code values as observed from

destructive testing of T-beam bridges due to the redundancy provided by beam flexural-shear and slab

punching-shear mechanisms that are both present (Al-Mahaidi et al. 2000, Song et al. 2002).

21

7 PRACTICAL EVALUATION OF THE T-BEAM POPULATION

We may rationalize ignoring many of the mechanisms that provide higher load capacity rating of RC T-

beam bridges for a need to be conservative. However, analytical sensitivity studies clearly indicate that the

actual lateral load distribution in a single-span RC T-beam bridge between various beams due to a truck

positioned on the deck may be considerably more effective than what is obtained by using the AASHTO

distribution factors. In addition, the distribution factors for RC T-beam bridges computed using LRFD

(AASHTO 1994) are determined to be more conservative than the LFD based distribution factors

(AASHTO 1989) shown in Table 4 and as illustrated in Figure 11. PennDOT bridge and district engineers

are favorable to incorporating the accurate extent of the live load distribution mechanism into load rating.

Therefore, it makes sense to examine and derive the equivalent distribution factors more accurately by

taking into account the actual geometry and detailing of the T-beam bridges. Although the AASHTO load

distribution factors were derived using FE analysis, it is clear that detailed 3D FE solid models that

represent the geometric characteristics of the population with a fine discretization are needed for

improved precision (Catbas et al, 2003). The distribution factors derived from analyses of geometric

replica 3D FE models that precisely represent PA’s T-beam bridge population will help improve load the

rating of these bridges while still strictly conforming to the AASHTO standards and provisions. The

equations that were derived for the Pennsylvania single span T-beam bridges are given in Table 5.

Therefore, 40 T-beam bridges representing the entire geometry and design spectrum of the RC T-beam

population in Pennsylvania were identified for deriving the lateral distribution factors more accurately

using 3D FE models. These models were constructed and analyzed under critical positions of two

simultaneous rating trucks. The diaphragm beams, parapets and boundary restraints were ignored in the

analyses as all beams were assumed to be simply-supported, permitting axial movement. The maximum

flexural and shear demands from FE analysis were compared to the corresponding demands obtained from

BAR7 analyses of the same bridge conducted by applying one-half of a truck as live load. The ratio of the

22

maximum demands from 3D FE analysis and BAR7 analysis for the same bridge provided an equivalent

lateral load distribution factor for that bridge.

These studies indicate that by using the distribution factors obtained from 3D microscopic FE models that

precisely represent the geometry of T-beam bridges, and by strictly complying with all of the capacity and

demand calculation requirements of AASHTO, it is still possible to increase the load rating of RC T-beam

bridges by 10%-55% depending on the geometry of the bridge. The distribution factors are expressed in

terms of simple equations in closed form and can be very easily implemented in load rating procedures.

23

8 CONCLUSIONS AND RECOMMENDATIONS

1) This study demonstrated that the statistical sampling strategy may serve as an effective approach for

condition assessment and management of large bridge populations that share common structural and

condition parameters. This strategy requires a determination of the critical nominal and as-is

condition parameters that govern the load capacity of a bridge population that will be evaluated by

statistical sampling.

2) This study took advantage of FE modeling and load testing in the context of structural identification

of a statistically representative sample of a bridge family to characterize the entire population. This

approach makes it possible to take maximum advantage of the bridge-type specific heuristics that has

been accumulated by experienced District engineers, and integrate this with the advanced

technological tools that offer reliable and measurement-based determination of serviceability and load

capacity.

3) Data on 27 RC T-beam bridges that were inspected and documented and the four bridges that were

subjected to controlled tests and structural-identification is summarized in a Report (Catbas et al.

2003). Results of in-depth analyses by field-calibrated 3D FE models for four of the test bridges

were adequate to describe and quantify the mechanisms that affect both the demand side and the

capacity side of the load rating equation. The studies revealed that load capacity rating of the RC T-

beam bridges by field-calibrated 3D FE models indicate rating factors that exceed the corresponding

factors obtained by BAR7 analysis by at least 2.5 times and up to five times. The actual load capacity

rating factors that would have been observed if the bridges were loaded to damage levels in the field

would in fact be much higher than even what is estimated by the FE analysis. The mechanisms that

lead to higher rating are consistent throughout the population, these are not temporary mechanisms,

and the reliability of their current existence has been verified by in-depth inspections of the 27 bridges.

4) Even if all of the mechanisms that are not typically included in the idealized modeling and analysis of

bridges by BAR7 are excluded, 3D FE models still indicate that it is possible to increase the load

24

rating of the population by between 10% and 55% due to only the enhanced live load distribution in

short single-span RC T-beam bridges. Alternative load distribution factors for RC T-beam Bridges in

Pennsylvania are developed for PADOT using 40 FE models that represent the entire population

geometry and design spectrum with a fine resolution. The corresponding equations and findings are

expected to impact the management of T-beam bridges after their review and approval by PADOT and

AASHTO. However, before we may take advantage of the more favorable distribution factors that

have been derived for PA’s single-span T-Beam population, it is recommended to perform carefully

designed field experiments on various decommissioned T-Beam bridges under controlled load levels

leading to damage and failure.

5) Whether shear or flexure governs the load capacity rating is one remaining very important issue since

this relates to the failure mode that should be expected in the case of overloading or the loss of

capacity due to continued deterioration and damage. The inherent deformability and resiliency

associated with a flexure-governed load rating is much greater than the corresponding attributes if

shear governs load capacity.

6) In an attempt to be conservative and to conform to AASHTO specifications, it is possible to exclude

the secondary mechanisms as in the case of BAR7 results. However, in reality the secondary

mechanisms do exist and they change the load demands within the structure and this may lead to a

shear failure mechanism governing the load rating. This study clearly illustrated that shear load rating

may govern due to existence of mechanisms that are ignored in an idealized simple-beam modeling.

Consequently, in spite of the apparent conservatism of the AASHTO provisions, shear may in fact

become the governing failure mode for many bridges operating daily under traffic as a result of

deterioration. The field studies, experiments and analyses here have indicated that the contributions

due to the secondary elements and mechanisms that enhanced load-rating were always greater than any

negative contributions due to existing deterioration and damage that reduce load rating. However, the

experiments were conducted under proof load levels (upper threshold of operating loads). Unless

25

testing is conducted at damage levels, a desired margin of safety may not be assured against

undesirable failure modes past the initiation of yielding, especially if extreme cases of deterioration

and damage are present. Tests should be conducted at higher load levels and up to failure to reveal the

extent of any adverse effects of existing deterioration and damage on load distribution, and help to

verify that the load distribution coefficients remain valid at higher load levels up to failure.

7) Therefore, it is recommended as prudent to evaluate the actual load capacity and failure modes of

several decommissioned T-Beam bridges by destructive testing. It is possible to conduct destructive

tests under loading by actuators reacting against rock-anchors, and with loading blocks that properly

simulate load distribution under a tandem-axle (Aktan et al, 1993). In addition, destructive testing

should be accompanied by nonlinear finite element analysis in order to derive maximum benefit from

them. A safe and meaningful design of the destructive testing in conjunction with nonlinear analysis

may permit the results to be generalized to the entire population and this may serve for validation of

the findings for the highest actual and the highest utilizable load rating of T-beam bridges.

8) The bridge management consequences of the conclusion reached in this study are not insignificant.

Currently, Pennsylvania has the third largest RC T-beam population after CA and KY and has the

most structurally deficient and functionally obsolete bridges T-beam bridges in the US (NBI, 2003).

Without a rational approach for taking advantage of their inherent capacity, greater numbers of these

bridges will soon have be posted and replaced. The financial impact of deferring the replacement of

the posted bridges for a decade is expected to be large amount of public funds. However, with the

promise of statistical sampling strategies integrated with objective measurements and advanced

analytical technologies, it would be possible to develop more effective inspection procedures, to

obtain a better estimate of the load rating of the bridge which will enable us to better evaluate, operate

and maintain infrastructure populations.

26

9 ACKNOWLEDGMENTS

The Pennsylvania Department of Transportation and the Federal Highway Administration have sponsored

the research reported here. The authors are grateful to Mr. Gary Hoffman, Mr. Scott Christie and Ms. Patti

Kiehl (PennDOT), and Drs. Steve Chase, Hamid Ghasemi and Mr. William Williams (FHWA) for their

support in making this project possible. Also, assistance and coordination of PennDOT District 6

engineers, especially Mr. Larry Ward during the conduct of the tests and field visits is appreciated. We

acknowledge the contributions of Messers. Grimmelsman, Barrish, von Haza Radlitz and Drs. Hasancebi

and Pervizpour (Drexel Infrastructure Institute) at different stages of the project. Especially, Dr. Oguzhan

Hasancebi’s invaluable contributions in developing and analyzing the finite element models are greatly

appreciated.

The contents of this paper reflect the views of the authors. The contents do not necessarily reflect the

official view or policies of the Commonwealth of Pennsylvania.

27

10 REFERENCES

AASHTO (1989), “Guide Specifications for Strength Evaluation of Existing Steel and Concrete Bridge,”

Washington D.C. 2001.

AASHTO (1994), “LRFD Bridge Design Specifications”, Washington D.C.

AASHTO (2000), “Manual for condition evaluation of bridges.” American Association of State Highway

and Transportation Officials, Washington D.C.

ACI (1998) Building Code Requirements for Structural Concrete, American Concrete Institute

Committee 318.

Aktan, A.E., Catbas, F.N., Grimmelsman, K.A. and Pervizpour, M. (2002), “Development of a Model

Health Monitoring Guide for Major Bridges”, Report, submitted to FHWA Research and

Development.

Aktan A.E., Catbas F. N., Turer A., Zhang Z.F. (1998), “Structural identification: Analytical aspects.”

Journal of Structural Engineering, ASCE, 124(7), 817-829.

Aktan A.E., Farhey D. N., Helmicki, A.J., Brown, D.L., Hunt, V.J., Lee, K.L., and Levi, A. (1997),

“Structural identification for condition assessment: experimental arts” Journal of Structural

Engineering, ASCE, 123(12), 1674-1684.

Aktan, A.E., Zwick, M.J., Miller, R.A. and Shahrooz, B.M., (1993) “Nondestructive and Destructive

Testing of a Decommissioned RC Slab Highway Bridge and Associated Analytical Studies,” paper

presented at TRB 92 and published in Transportation Research Record (TRR) 1371, pp:142-153;

National Academy Press, Washington, D.C. 1993.

28

Al-Mahaidi, R., Taplin, G. and Giufre, A. (2000) “Load Distribution and Shear Strength Evaluation of an

Old Concrete T-beam Bridge,” Transportation Research Record - Journal of the Transportation

Research Board 1696, Vol.1, pp.52-61.

Ang, A. A-S. and Tang, W. H., (1975) “Probability Concepts in Engineering Planning and Design,

Volume I - Basic Principles,” John Wiley & Sons, New York, 1975.

ArcVIEW Software (2001), Environmental Systems Research Institute, Inc.

ASCE (2003) Report Card for America’s Infrastructure, http://www.asce.org/reportcard/

Catbas, F.N., Ciloglu, S.K. Hasancebi, O. and Aktan, A.E. (2002) "Fleet Strategies for Condition

Assessment and Its Application for Re-qualification of Pennsylvania's Aged T-beam Bridges", Paper

No.02-3890, presented at the 81th Annual Meeting of Transportation Research Board and

submitted for publication in TRR, Washington DC, 2002.

Catbas, F. N., Hasancebi, O., Ciloglu, S. K., and Aktan, A.E. (2003) “Re-qualification of Aged

Reinforced Concrete T-beam Bridges in Pennsylvania,” Project Report Submitted to Pennsylvania

Department of Transportation Bureau of Design, Bridge Quality Assurance Division by Drexel

Intelligent Infrastructure and Transportation Safety Institute, Drexel University.

Chase, S.B (2001) “Hi-tech Inspection,” Civil Engineering Magazine, ASCE, pp.62-65, September.

FHWA (1995), Recording and Coding Guide for the Structure Inventory an Appraisal of the Nation’s

Bridges, U.S. Department of Transportation, Federal Highway Administration, Report Number

FHWA-PD-96-001, Washington D.C., December 1995.

FHWA (2001), “Reliability of Visual Inspection for Highway Bridges," FHWA Report Nos. FHWA-

RD-01-020 and FHWA-RD-01-021, 2001.

29

FHWA (2004), Non-Destructive Evaluation Validation Center,

http://www.tfhrc.gov/hnr20/nde/home.htm, 2004

Huria, V., Lee, K.L. and Aktan, A.E. (1994),”Different Approaches to Rating Slab Bridges”, Technical

Note, Journal of Structural Engineering, ASCE, October 1994.

Livingston, R.A. and Amde, A.M. (2000), "Nondestructive Test Field Survey for Assessing the Extent of

Ettringite-Related Damage in Concrete Bridges," presented in International Symposium on the

Nondestructive Characterization of Materials, Karuizawa, Japan.

MacGregor, J.G. (1998) “Reinforced Concrete-Mechanics and Design”, Prentice Hall.

Madanat, S.M., Karlaftis, M.G. and McCarthy, P.S. (1997), “Probabilistic Infrastructure Deterioration

Models with Panel Data” ASCE J. of Infrastructure Systems, Vol. 3, No. 1, pp.4-9.

New York Times (2002), National Briefing in Washington, December 7, 2002.

NBI, (1998), National Bridge Inventory Data (NBID).Rep.No.FHWA-PD-96-001, Washington D.C.,

1998.

NBI, (2003), National Bridge Inventory Data (NBID). http://www.fhwa.dot.gov/bridge/britab.htm,

Washington D.C.

Park, T. and Gamble, W., (2000) “Reinforced Concrete Slabs”, John Wiley and Sons, 2nd Edition.

PennDOT (2001), Bridge Analysis and Rating Program - BAR7, V7.10, PennDOT Bureau of Information

Systems Application Development Division.

Roberts, J. and Shepard, R. (2001), "Bridge Management for the 21st Century", Proceedings of the SPIE's

6th International Symposium on NDE and Health Monitoring and Diagnostics, Vol. 4337, pp.48-59,

Newport Beach, CA.

30

Shahrooz, B., Ho, I.K., Lee, K.-L., Aktan, A.E., de Borst, R., Blaauwendraad, J., van der Veen, C., Iding,

R. H., Miller, R. A. (1994), “Nonlinear Finite Element Analysis of Deteriorated RC Slab Bridge,”

Journal of Structural Engineering, Vol. 120, No. 2, pp. 422-440, ASCE.

Song, H.W., You, D.W., Byun, K.J. and Maekawa, K. (2002) “Finite Element Failure Analysis of

Reinforced Concrete T-girder Bridges,” Engineering Structures Vol.24, pp.151-162.

"Standards For Old Bridges" (1983) Commonwealth of Pennsylvania, Department of Transportation,

Bureau of Highway Design Bridge Division (from 1931 to 1940) Volumes 1-2, May 1983.

31

11 LIST OF FIGURES

Figure 1: Single Span T-Beam Population (grey squares) and the Locations of Statistically Representative

sample of 60 Bridges (dark squares).

Figure 2: Critical Parameters of the Entire Single Span T-Beam Bridge Population and the Statistically

Representative sample of 60 Bridges

Figure 3: Swan Road Bridge: General and Close-up Views

Figure 4: 3D Finite Element Modeling Of the T-Beam Bridges Using Solid and Frame Elements

Figure 5: Swan Road Bridge Field Testing and Instrumentation Plan

Figure 6: Step-By-Step Calibration of the Swan Bridge FE Model to Match Test Data

Figure 7: Extreme Deterioration/Damage Simulation

Figure 8.a: Swan Road Bridge Calibrated FEM Load Rating. b: Swan Road Bridge Damage Simulation

Load Rating

Figure 9: Swan Road Bridge Rating Factor Parameter Sensitivity Study.

Figure 10: Tested Bridges.

Figure 11: Live Load Moment Distribution Factors as a Function of Span Length for Bridges with No

Skew

12 LIST OF TABLES

Table 1: Summary of Load Test Results

Table 2: Summary of Dynamic Test and FWD Results

Table 3: Correlation of Modal Frequencies for Swan Road Bridge

Table 4: Current AASHTO Equations for T-beam Bridges

Table 5: Equations Derived for Single Span Bridge Population

32

Table 1: Summary of Load Test Results

Truck Load Applied

(2)

Max. Deflection

(3)

L/800

(4)

Max. Rebar Stress

(5)

Max. Concrete Stress

(6) Bridge Name

(1) (kip) (in) (in) (psi) (psi)

Swan 98 0.015 0.032 886 120

1 kip=4.45 kNt; 1 in=2.54 cm; 1 psi=6.89 kPa

33

Table 2: Summary of Dynamic Test and FWD Results

Modal Frequencies (Hz) (2)

Flexibility Coefficients (in/kip x 10-3) (3) Bridge Name

(1) Mode 1 Mode 2 Mode 3 Load Test Impact Test FWD

Swan 22.38 41.26 55.40 0.409 0.415 0.525

1 kip=4.45 kNt; 1 in=2.54 cm

34

Table 3: Correlation of Modal Frequencies for Swan Road Bridge

Mode No (1)

Test (Hz) (2)

Preliminary (Hz) (3)

Calibrated Model (Hz) (4)

1 22.38 14.64 25.83

2 41.26 27.31 35.69

3 55.40 34.19 39.43

35

Table 4: Current AASHTO Equations for T-beam Bridges

S (beam spacing), L (span of beam), Kg (long. Stiffness parameter), ts (slab thickess)

(See AASHTO LRFD Specs for details)

AASHTO LRFD Bridge Design Specs.

if θ <30 then c1=0.0if θ >60 then θ =60

AASHTO Standard Specs for Highway Bridges (LFD)

If S exceeds 10 ft: Assume flooring between stringers acts as a simple beam with the load on each stringer being the wheel load reaction

Moment DF for Two Design Lane Loaded

Range

AASHTO LRFD Bridge Design Specs.

if θ <30 then c1=0.0if θ >60 then θ =60

AASHTO Standard Specs for Highway Bridges (LFD)

If S exceeds 10 ft: Assume flooring between stringers acts as a simple beam with the load on each stringer being the wheel load reaction

Moment DF for Two Design Lane Loaded

Range

( )[ ]5.113

2.06.0

tan10.125.9

075.0 θcLt

K

LSS

gs

g −

+=

424020125.4165.3

≥≤≤≤≤≤≤

b

s

NLtS

AASHTO LRFD Bridge Design Specs.

AASHTO Standard Specs for Highway Bridges (LFD)For T-beams S < 6 ft

Shear DF for Two Design Lane LoadedRange

AASHTO LRFD Bridge Design Specs.

AASHTO Standard Specs for Highway Bridges (LFD)For T-beams S < 6 ft

Shear DF for Two Design Lane LoadedRange

+

+= θtan

0.1220.01

35122.0

3.032

g

s

KLtSS

g

600

710

4

24020

125.4165.3

≤≤

≤≤≥

≤≤≤≤≤≤

θ

MKK

N

L

tS

g

b

s

621 S

g ∗=

+∗=S

Sg

41

21

=

5.025.0

31 1225.0

LS

Lt

Kc

s

g

36

Table 5: Equations Derived for Single Span Bridge Population

g=distribution factor, L=clear span as given in PA Standards for Old Bridges, θ=skew angle

32’-42’θ=30-45

24’-32’θ=30-45

32’-42’θ=0-30

24’-32’θ=0-30

Moment DF for Two Design Lane Loaded (for PA T-beams given in Standards for Old Bridges)

Range

32’-42’θ=30-45

24’-32’θ=30-45

32’-42’θ=0-30

24’-32’θ=0-30

Moment DF for Two Design Lane Loaded (for PA T-beams given in Standards for Old Bridges)

Range

32’-42’θ=0-45

24’-32’θ=0-45

Shear DF for Two Design Lane Loaded (for PA T-beams given in Standards for Old Bridges)

Range

32’-42’θ=0-45

24’-32’θ=0-45

Shear DF for Two Design Lane Loaded (for PA T-beams given in Standards for Old Bridges)

Range

[ ]

−∗++∗∗−∗= −

15101.0185.1106170115 52 θ

LLg

[ ]

−∗+∗+∗−= −

151013.0104007888.62 5 θ

Lg

[ ]

−∗++∗∗−∗= −

152

5021.0009.110506745.94 52 θLLg

[ ]

−∗+∗+∗= −

152502.0103347606.36 5 θLg

[ ]

−∗++∗∗−∗= − 1

452

022.018664.2103.91455.124 52 θLLg

[ ]

−∗+∗+∗= − 1

452

032.01030315744 5 θLg

Figure 1: Single Span T-Beam Population And The Locations Of Statistically Representative 60 Bridges

Entire Population Sample Population

Figure 2: Critical Parameters Of The Entire Single Span T-Beam Bridge Population And The Statistically Representative 60 Bridges

Condition Rating

5

36%

623%

7 to 818% 4

20%

33%

Age

1929 to 1938

34%

< 192924%

>1948

24%

1939 to 1948

18%

Skew Angle(deg)> 50

1%

8 to 2219%

23 to 3722%

38 to 5020%

0 to 738%

> 55 ft 0%

16 ft to 32 ft

64%33ft to 40 ft

22%

41 ft to 55 ft14%

Span

Entire T-Beam Bridge Population

> 500%

0 to 7

43%8 to 22

18%

23 to 3722%

38 to 50

17%

Skew Angle(deg)

Skew Angle

(degrees)

> 55 ft 0%

16 ft to 32 ft

62%33ft to 40ft18%

41 ft to 55 ft

20%

Span

< 1929

30%

1929 to 1938

30%

1939 to 194817%

> 1948

23%

Age

3

7%7 to 8

15%6

17%

4

27%

5

34%

Condition Rating

Statistical Representative 60 T-Beam Bridges

Condition Rating

5

36%

623%

7 to 818% 4

20%

33%

Age

1929 to 1938

34%

< 192924%

>1948

24%

1939 to 1948

18%

Condition Rating

5

36%

623%

7 to 818% 4

20%

33%

Condition Rating

5

36%

623%

7 to 818% 4

20%

33%

Age

1929 to 1938

34%

< 192924%

>1948

24%

1939 to 1948

18%

Age

1929 to 1938

34%

< 192924%

>1948

24%

1939 to 1948

18%

Skew Angle(deg)> 50

1%

8 to 2219%

23 to 3722%

38 to 5020%

0 to 738%

> 55 ft 0%

16 ft to 32 ft

64%33ft to 40 ft

22%

41 ft to 55 ft14%

Span Skew Angle(deg)> 50

1%

8 to 2219%

23 to 3722%

38 to 5020%

0 to 738%

Skew Angle(deg)> 50

1%

8 to 2219%

23 to 3722%

38 to 5020%

0 to 738%

> 55 ft 0%

16 ft to 32 ft

64%33ft to 40 ft

22%

41 ft to 55 ft14%

Span

> 55 ft 0%

16 ft to 32 ft

64%33ft to 40 ft

22%

41 ft to 55 ft14%

Span

Entire T-Beam Bridge Population

> 500%

0 to 7

43%8 to 22

18%

23 to 3722%

38 to 50

17%

Skew Angle(deg)

Skew Angle

(degrees)

> 55 ft 0%

16 ft to 32 ft

62%33ft to 40ft18%

41 ft to 55 ft

20%

Span> 50

0%

0 to 7

43%8 to 22

18%

23 to 3722%

38 to 50

17%

Skew Angle(deg)> 50

0%

0 to 7

43%8 to 22

18%

23 to 3722%

38 to 50

17%

Skew Angle(deg)

Skew Angle

(degrees)

> 55 ft 0%

16 ft to 32 ft

62%33ft to 40ft18%

41 ft to 55 ft

20%

Span

Skew Angle

(degrees)

> 55 ft 0%

16 ft to 32 ft

62%33ft to 40ft18%

41 ft to 55 ft

20%

Span

< 1929

30%

1929 to 1938

30%

1939 to 194817%

> 1948

23%

Age

3

7%7 to 8

15%6

17%

4

27%

5

34%

Condition Rating

< 1929

30%

1929 to 1938

30%

1939 to 194817%

> 1948

23%

Age

< 1929

30%

1929 to 1938

30%

1939 to 194817%

> 1948

23%

Age

3

7%7 to 8

15%6

17%

4

27%

5

34%

Condition Rating

3

7%7 to 8

15%6

17%

4

27%

5

34%

Condition Rating

Statistical Representative 60 T-Beam Bridges

Figure 3: Swan Road Bridge: General And Close-Up Views

Cross Section of the Model

16.85”

15.5”

15.75”8.5”

T-Beams

Parapet

End Diaphragm

Rebar Layout Structural Details and Boundary Conditions

Statistics of The ModelNumber of DOF =108243Number of Solid Elements = 22940Number of Frame Elements = 7636

3.375”

12”3.375”

Figure 4: 3D Finite Element Modeling Of The T-Beam Bridges Using Solid And Frame Elements

1 in=2.54 cm

Displacement Sensor Location

Steel Strain Sensor Location Concrete Strain Gauge Location

A-A

A B C D E F

3

2

1

CL

CL

B-B

Truck and Sensor Locations:

Figure 5: Swan Road Bridge Field Testing And Instrumentation Plan

Figure 6: Step By Step Calibration Of The Swan Bridge FE Model To Match Test Data

b) Local Calibration and Correlation

Transverse Centerline Deflection of the Superstructure (Test vs. Models)

Def

lect

ion

(in

)

-0.010

-0.020

-0.030

0

-0.040

-0.050

-0.060

-0.070

Section A-A

A2 B2 C2 D2 E2 F2

-0.010

-0.020

-0.030

0

Def

lect

ion

(in)

-0.040

-0.050

-0.060

-0.070

Deflection of the T-Beam "C" (Test vs. Models)

Superstructure

C3 C2 C1

Section B-BLoad Test Truck

51.5 kips 48.0 kips

a) Regional Calibration and Correlation

Ste

el S

tres

s (p

si)

300

900

0

1200

1500

1800

2100

2500

600

Transverse Centerline Steel Rebar Stresses (Test vs. Models)

Co

ncr

ete

Str

ess

(psi

)

0

40

80

120

160

200 Concrete Stressesalong T-beam "C" (Test vs. Models)

A2 B2 C2 D2 E2 C2C3 C1

Models for Calibration –from Nominal to Calibrated

1 in=2.54 cm1 psi = 6.9 kPa

Assumed Damage: 40% concrete spalling at all beams. 50% rebar corrosion at lower layer and stirrups and 20% at upper layer rebars

Deterioration/Damage Simulation

d=24'' (Swan)d=28.5'' (Manoa)

40% of the entire depth

8.5"

~3.4"~2.6"

Figure 7: Extreme Deterioration/Damage Simulation

As-is Condition with All Elements, End RestraintsPin-pin Supports (using calibrated FEM)

Swan Road Bridge; RFM = 3.18 RFV = 2.69

(b)

Figure 8: Swan Road Bridge a) Calibrated FEM Load Rating, b) Damage Simulation Load Rating

As-is Condition with All Elements, End RestraintsPin-pin Supports (using calibrated FEM)Projected Extreme DeteriorationSwan Road Bridge; RFM = 2.11 RFV = 2.30

(a)

Pin-roller Supports w/ Parapetsw/o Diaphragmsw/o Pavement Thrust

RFM = 1.88 RFV = 3.10

x DF

Pin-roller Supports w/o Parapetsw/o Diaphragmsw/o Pavement Thrust

RFM = 1.44 RFV = 2.64

Pin-roller Supports w/o Parapetsw/o Diaphragmsw/o Pavement Thrustw/ Extreme Deterioration

RFM = 0.88 RFV = 1.97

Pin-Pin Supports w/ Parapetsw/ Diaphragmsw/o Pavement Thrust

RFM = 3.05 RFV = 3.54

Pin-Roller Supports w/ Parapetsw/ Diaphragmsw/o Pavement Thrust

RFM = 1.99 RFV = 3.90

BAR7 AnalysisRFM = 1.27 RFV = 1.80

Figure 9(a)-(f): Swan Road Bridge Rating Factor Parameter Sensitivity Study

(a) (b) (c)

(d) (e) (f)

Ratio (FEM/AASHTO) = 5.10

BAR7 Rating = 1.01Field Calibrated FEM = 5.15

Ratio (FEM/AASHTO) = 2.75

BAR7 Rating = 1.22Field Calibrated FEM = 3.35

Ratio (FEM/AASHTO) = 2.50

BAR7 Rating = 1.27Field Calibrated FEM = 3.18

Ratio (FEM/AASHTO) = 3.46

BAR7 Rating = 0.92Field Calibrated FEM = 3.18

Figure 10: Tested Bridges

Live Load Moment Distribution Factors for 90 deg Skewed Bridges

0.200

0.300

0.400

0.500

0.600

0.700

0.800

22 24 26 28 30 32 34 36 38 40 42 44

Span Length (ft)

Dis

trib

uti

on

Fac

tors

Drexel FEM Moment Distribution Factor for 90 deg SkewCurve Fit and Closed Form DF Formulation for Two Lanes

AASHTO LFD Moment Distribution FactorAASHTO LRFD One Lane Moment Distribution Factor for 90 deg Skew

AASHTO LRFD Two Lane Moment Distribution Factor for 90 deg Skew

Figure 11: Live Load Moment Distribution Factors As A Function Of Span Length For Bridges With No Skew

1 ft=0.305 m