Sie sind auf Seite 1von 277

Verification of ODOT’s Load Rating

Analysis Programs for Metal Pipe and Arch


Culverts

Halil Sezen (PI)


Patrick P. Fox (co-PI)
Kyong Y. Yeau

Prepared in cooperation with


the Ohio Department of Transportation,

and the U.S. Department Transportation,


Federal Highway Administration

State Job Number 134225

August 2009
1. Report No. 2. Government Accession No. 3. Recipient’s Catalog No.
FHWA/OH-2009/6
4. Title and subtitle 5. Report Date
August 2009
VERIFICATION OF ODOT’S LOAD RATING ANALYSIS 6. Performing Organization Code
PROGRAMS FOR METAL PIPE AND ARCH CULVERTS
7. Author(s) 8. Performing Organization Report
No.
Halil Sezen, Patrick P. Fox, and Kyong Y. Yeau
10. Work Unit No. (TRAIS)

9. Performing Organization Name and Address 11. Contract or Grant No.


The Ohio State University State Job No. 134225
Department of Civil and Environmental Engineering and Geodetic Science 13. Type of Report and Period
470 Hitchcock Hall Covered
2070 Neil Avenue Final Report
Columbus OH 43210
12. Sponsoring Agency Name and Address 14. Sponsoring Agency Code
Ohio Department of Transportation
1980 West Broad Street
Columbus, OH 43223
15. Supplementary Notes
Prepared in cooperation with the Ohio Department of Transportation (ODOT)
16. Abstract
The main objective of this study was to evaluate and improve ODOT’s current load rating procedures for corrugated
metal culverts. This objective is achieved by testing 39 in-service culverts under static and dynamic loads, by
evaluating the response of test culverts using available theoretical methods and numerical simulations, and by
evaluating and advancing the current analysis tools and load rating methods based on the analytical and
experimental evidence generated in this research.

The experimental program was conducted to investigate the influence of several parameters on the field performance
of culverts in Ohio. These parameters include backfill height, various static and dynamic load applications, and
existing condition, size, shape, and other properties of corrugated metal culverts. Experimental results show that
culvert deflections decrease nonlinearly with increasing backfill height. Deflections and strains were nearly zero in
deep culverts with backfill height larger than 13 ft (4 m). Under static and dynamic truck loading, culvert deflections
and strains increased significantly when the backfill height was less than about 6.5 ft (2.0 m). Responses of some of
the test culverts were simulated using the two-dimensional finite element program CANDE. Deflections predicted
from CANDE analysis were larger than deflections measured in the field. However, the moments and thrusts
calculated from experimental strains were similar to those calculated using theoretical methods and CANDE.

Experimental data and available theoretical studies were used to evaluate the current load rating methods.
Recommendations are made to improve the analysis and evaluation procedures for corrugated metal culverts.
Recommendations are based on an extensive review of load rating procedures and design practices, experimental
data from 39 test culverts, and theoretical and numerical investigations. The recommended load rating procedures do
not consider the effect of cover depth. New capacity reduction factors are introduced for culvert wall and seam,
which require different appraisals for wall and seam during annual inspections. We recommend changes to rating
factors for deep culverts and culverts subjected to low live load stresses.
17. Key Words 18. Distribution Statement
Load rating, rating factor, corrugated metal culvert, culvert No restrictions. This document is available to
experiment, culvert behavior, deflection, thrust force, seam the public through the National Technical
strength, buckling
Information Service, Springfield, Virginia 22161
19. Security Classif. (of this report) 20. Security Classif. (of this page) 21. No. of Pages 22. Price
Unclassified Unclassified 263
Form DOT F 1700.7 (8-72) Reproduction of completed pages authorized
Verification of ODOT’s Load Rating
Analysis Programs for Metal Pipe and Arch
Culverts
Halil Sezen, Ph.D.
Associate Professor, Dept. of Civil and Environmental Engineering and Geodetic Science,
The Ohio State University, Columbus, OH 43210

Patrick P Fox, Ph.D.


Professor, Dept. of Civil and Environmental Engineering and Geodetic Science,
The Ohio State University, Columbus, OH 43210

Kyong Y. Yeau
Ph.D student, Dept. of Civil and Environmental Engineering and Geodetic Science,
The Ohio State University, Columbus, OH 43210

August, 2009

A Report to Sponsors:
The Ohio Department of Transportation,
and
the U.S. Department of Transportation,
Federal Highway Administration

State Job No. 134225

Prepared in cooperation with the Ohio Department of Transportation and


the U.S. Department of Transportation, Federal Highway Administration

The contents of this report reflect the views of the authors who are responsible for the facts and
the accuracy of the data presented herein. The contents do not necessarily reflect the official views
or policies of the Ohio Department of Transportation or the Federal Highway Administration.
This report does not constitute a standard, specification or regulation.
TABLE OF CONTENTS

Chapter Page

1. INTRODUCTION ……………………………………………………………… 1
1.1 Introduction and problem statement……………………………………….... 1
1.2 Objectives………………………………………………………………….... 4
1.3 Scope and organization…………………………………………………..…. 5

2. BACKGROUND INFORMATION AND PREVIOUS RESEARCH………..… 7


2.1 Experimental studies……………………………………………………....… 7
2.1.1 Culvert deflections……………………………………………………. 7
2.1.2 Corrugated plate forces……………………………………………….. 8
2.1.3 Buckling……………………………………………………………… 8
2.2 Analytical studies…………………………………………………………… 9
2.3 Theoretical studies…………………………………………………………... 9
2.3.1 Soil-structure interaction studies……………………………………... 10
2.3.2 Marston-Spangler method……………………………………………. 10
2.3.3 Ring compression theory…………………………………………...… 12
2.3.4 Buckling considerations……………………………………………… 14
2.3.5 Live loads…………………………………………………………..… 15
2.3.5.1 Boussinesq equations………………………………………….. 15
2.3.5.2 Load distribution assumptions………………………………… 16

3. EXPERIMENTAL RESEARCH AND CULVERT INFORMATION………… 18


3.1 Introduction……………………………………………………………….… 18
3.2 Instrumentation……………………………………………………………… 18
3.2.1 Deflections………………………………………………………….… 18
3.2.2 Strains………………………………………………………………… 19
3.3 Live load application……………………………………………………...… 21
3.4 Description of test culverts………………………………………………..… 24

4. EXPERIMENTAL TEST RESULTS………………………………………...… 46


4.1 Introduction……………………………………………………………….… 46
4.2 Measurement of culvert deformations…………………………………….… 49
4.2.1 Static response..………………………………………………………. 49
4.2.2 Dynamic response………………………………………………….… 55
4.3 Measurement of culvert strains…………………………………………...… 55
4.3.1 Static response……………………………………………………...… 58

i
4.3.2 Variation in strain gauge locations in different culverts…………...… 59

5. EVALUATION TEST RESULTS……………………………………………...… 67


5.1 Introduction……………………………………………………………….… 67
5.2 Relationship between the factors affecting culvert behavior ……………..… 67
5.3 Equivalent line load……………………………………………………….… 72
5.4 The effect of truck speed………………………………………………….… 74
5.5 Internal forces in plates……………………………………………………… 77

6. NUMERICAL MODELING AND ANALYSIS OF CULVERTS……………... 82


6.1 Introduction……………………………………………………………….… 82
6.2 CANDE program and modeling.………………………………………….… 82
6.3 Finite element models of corrugated metal culverts………………………… 85
6.3.1 Culvert modeling…………………………………..……………….… 85
6.3.2 Soil and pavement modeling……………………………………….… 86
6.6.3 Simulation of live loads………………………………………………. 88
6.4 Finite element analysis results…………………………………………….… 89

7. ODOT’s LOAD RATING PROCEDURE……………….………………….….. 95


7.1 Introduction……………….………………….………………….…………... 95
7.2 ODOT’s CMP-Excel program……………….………………….…………... 96
7.3 Basic information needed for load rating evaluation……………….……….. 98
7.4 Physical strength calculations……………….………………….…………… 100
7.4.1 Seam strength……………….………………….………………….….. 102
7.4.2 Wall strength……………….………………….………………….…... 102
7.4.3 Buckling strength……………….………………….…………………. 103
7.5 Thrust forces……………….………………….………………….…………. 104
7.5.1 Thrust capacity of wall……………….………………….…………… 104
7.5.2 Wall thrust due to dead load……………….………………….……… 104
7.5.3 Wall thrust due to live load plus impact……………….……………... 105
7.6 Rating factors for ring compression structures……………….……………... 106
7.6.1 Operating load rating……………….………………….……………... 108
7.6.2 Inventory load rating……………….………………….……………… 110
7.7 Discussion of ODOT’s load rating method……………….………………… 110
7.7.1 Change in gauge number……………….………………….…………. 112
7.7.2 Capacity of wall strength, Tcap………….………………….…………. 113
7.7.3 Rating factor based on minimum cover……………….……………… 118
7.7.4 Rating factor based on wall strength……………….………………… 119
7.7.5 Controlling rating factor……………….………………….………….. 120

8. PROPOSED LOAD RATING PROCEDURE……………….………………… 125


8.1 Introduction……………….………………….………………….…………... 125
8.2 Basic features of the proposed RF……………….………………….………. 126
8.3 Adequacy of the design gauge number for the cover depth………………… 128
8.3.1 Minimum soil cover……………….………………….………………. 129
8.3.2 The critical pressure……………….………………….………………. 130

ii
8.3.3 Proposed minimum cover……………….………………….………… 131
8.4 Thrust capacity of wall……………….………………….………………….. 139
8.4.1 General appraisal numbers for seam strength and wall area…………. 140
8.4.2 Reduction rate for seam strength and wall area……………….……… 148
8.4.2.1 Seam strength reduction rate……………….………………….. 148
8.4.2.2 Area reduction rate……………….………………….………… 150
8.4.3 Reduction factors for deflection and local buckling……………….…. 153
8.4.4 Comparison the proposed thrusts to current ODOT thrusts………….. 153
8.5 Proposed load rating……………….………………….………………….…. 154
8.5.1 Live load effect on the crown……………….………………….…….. 154
8.5.2 Operating and inventory load rating procedures……………….…….. 157
8.6 Application of proposed load rating……………….………………….…….. 158
8.7 Comparison current ODOT’s load rating to the proposed load rating……… 159
8.8 Load rating procedures based on AASHTO LRFD Specifications…………. 160
8.8.1 Design parameters……………….………………….………………... 163
8.8.2 Pressure on the crown due to live load……………….………………. 165

9. SUMMARY AND CONCLUSIONS……………….………………….………… 171

LIST OF REFERENCES…………….………………….……………….…………... 176

APPENDIX
A. EXPERIMENTAL TEST RESULTS FOR DEFLECTIONS.…………………. 183
B. EXPERIMENTAL TEST RESULTS FOR STRAINS……..………………….. 219
C. DUNCAN’S HYPERBOLIC SOIL MODEL…………………………………... 254
C.1 Introduction………………………………………………………………… 255
C.2 Hyperbolic stress-strain relationships……………………………………… 255
C.3 Theoretical development…………………………………………………… 256
D. CONDITION ASSESSMENT GUIDELINE FOR CORRUGATED METAL
CULVERTS…………………………………………………………………….. 259
D.1 Introduction………………………………………………………………… 260
D.2 Condition assessment………………………………………………………. 260

iii
LIST OF TABLES

Table Page

3.1 HS20-44 and ODOT BUSRET truck loads………………………………………. 22

3.2 Characteristics of corrugated metal culverts at 39 test sites..…………………….. 26

4.1 Applied truck loads and maximum measured displacements and strains. ……….. 47

5.1 Calculated stresses based on the plane-strain theory.…………………………….. 80

5.2 Maximum bending moment and axial thrust in the 39 test culverts……………… 81

6.1 Soil parameters used for the hyperbolic soil model in culvert analysis…..……… 88

6.2 Measured, predicted (CANDE), and theoretical (Equation 2.1) deflections at the
crown……………………………………………………………………………... 92

6.3 Measured, predicted, and theoretical thrust at the crown………………………… 93

7.1 Sectional properties of 6 × 2 in. (152 × 51 mm) corrugation plate (AISI, 1994,
Table 2.10)……..…………………………………………………………………. 101

7.2 Minimum longitudinal seam strengths for 6 × 2 in. (152 × 51 mm) corrugation
plate (Section 12.6.2 of AASHTO, 2002)………………………………………... 103

7.3 Maximum allowable moment strength (Duncan and Drawsky, 1983)…………… 110

7.4 Comparison of initial design and current gauge numbers………………………... 114

7.5 Design and current wall strength capacity for 39 culverts……………………….. 116

7.6 Design and current rating factors for 39 test culverts…………………………….. 122

7.7 Controlling RF for the worst case………………………………………………… 123

8.1 Pipe-arch size and layout details (18 in. (457 mm) corner radius) constructed
with 6×2 in. (152 × 51 mm) corrugated structural plate (AISI,1994)…………..... 133

iv
8.2 The minimum covers in feet for the standard culverts with different gauge
numbers.…………………………………………………………………………... 136

8.3 The proposed minimum covers in feet for the standard culverts with different
gauge numbers………………………………………………………………......... 138

8.4 Culvert inspection reports for the plate areas of 39 test culverts…………………. 141

8.5 Culvert inspection reports for the seams of 39 test culverts……………………… 145

8.6 Proposed seam strength reduction rate for corrugated metal pipe arches………... 149

8.7 Area reduction rate for corrugated metal pipe arches…………………………….. 152

8.8 Comparison of the proposed thrusts to current ODOT thrusts.…………............... 155

8.9 Comparison of the proposed RF to current ODOT RF…………………………… 161

8.10 Controlling RF for the worst case………………………………………………… 162

8.11 Multiple presence factor (AASHTO LRFD Table 3.6.1.1.2-1)…………………... 164

8.12 The factors used in LRFD RF calculations for the 39 test culverts………………. 169

8.13 Thrust due to live load and current rating factors for the 39 test culverts………... 170

v
LIST OF FIGURES

Figure Page

1.1 Typical large-span culvert shapes.………………………………………………... 2

2.1 Pressure distribution assumed in the Marson-Spangler method. ………………… 11

2.2 Pressure distribution assumed by the ring compression theory for a pipe culvert.. 13

2.3 Pressure distribution assumed by the ring compression theory for a pipe arch
culvert…………………………………………………………………………….. 14

2.4 Vertical stress caused by a point load.……………………………………………. 16

2.5 Distributed load area for single dual wheel of HS-20 truck……………………… 17

3.1 a) instrumentation frame, and b) displacement sensor locations…………………. 20

3.2 Strain gauge locations (circled channels indicate a strain gauge placed in
longitudinal direction of culvert). ……………………………………………...… 22

3.3 Test truck positions for ten static load cases……………………………………... 23

3.4 Corrugated metal culvert locations tested………………………………………… 25

3.5 The inspected condition of the culverts prior to testing.……………………….… 30

3.6 The inspected condition of the culverts before testing…………………………… 34

3.7 The inspected condition of the culverts before testing…………………………… 37

3.8 The inspected condition of the culverts before testing…………………………… 41

3.9 The inspected condition of the culverts before testing…………………………… 45

4.1 Deflections in culvert MRW-019-0456 under static and dynamic loads………… 51

4.2 Deflections in culvert MAD-038-0935 under static and dynamic loads…………. 52

vi
4.3 Deflections in culvert JAC-327-0077 under static and dynamic loads…………... 53

4.4 Deformed shape and maximum deflection of culvert MAD-038-0935 for ten
static load cases (deflections are magnified 254 times)………………………….. 54

4.5 Deflections of JAC-327-0077 for truck speeds varying from 5 to 40 mph (8 to 64


km/hr)…………………………………………………………………………….. 56

4.6 Measured strain versus time plot for culvert MAD-038-0935…………………… 60

4.7 Measured strain versus time plot for culvert MRW-095-1796.………………….. 62

4.8 Measured strain versus time plot for culvert JAC-327-0077…………………….. 64

4.9 Strain gauge locations of the culvert LIC-079-2276……………………………... 66

5.1 Maximum culvert deflection versus culvert span length…………………………. 69

5.2 Maximum culvert deflection versus culvert height………………………………. 69

5.3 Measured response for 40 corrugated metal culverts: (a) maximum deflection
and (b) maximum compressive strain versus backfill height…………………….. 70

5.4 Maximum culvert deflection versus total truck weight…………………………... 71

5.5 Maximum culvert deflection versus general appraisal number…………………... 71

5.6 Idealization of a point load to a line load (Katona et al., 1976)………………….. 73

5.7 Equivalent load distribution due to wheel loads over a culvert………………….. 75

5.8 Maximum culvert deflection versus AASHTO equivalent line load…………….. 75

5.9 Normalized dynamic loading time ratio versus truck speed……………………… 76

5.10 Maximum deflection ratio versus truck speed (Δ denotes deflection)…………... 76

5.11 Plane stress state in a cross section of corrugated plate………………………….. 78

5.12 Geometry to perform data analysis for culvert forces…………………………… 79

6.1 Half culvert model with 84, 183, and 339 nodes…………………………………. 84

6.2 Effect of mesh size on analysis results…………………………………………… 84

vii
6.3 Bilinear stress-strain relationship used for metal culvert models………………… 85

6.4 Beam elements used to model a corrugated metal culvert…………..…………… 86

6.5 Finite element model for corrugated metal culverts……………………………… 87

6.6 Deformed shapes of MAD-38-0935 depending on load application positions…... 91

6.7 Maximum deflections obtained from experiments, analyses, and theoretical


calculations……………………………………………………………………….. 94

6.8 Measured, predicted, and theoretical thrust at the crown………………………… 94

7.1 ODOT load rating flow chart…………………………………………………….. 97

7.2 Notation relating to culvert geometry…………………………………………….. 98

7.3 Reduction factor for buckling strength due to deflection in round pipes (NCSPA,
1995)……………………………………………………………………………… 99

7.4 Determination of actual culvert radius by field measurement (NCSPA, 1995)...... 100

7.5 Geometry of 6 × 2 in. (152 × 51 mm) corrugation plate…………………………. 101

7.6 Prism load and live load surcharge over culvert………………………………….. 105

7.7 HS-20 load distribution through soil cover over culvert in longitudinal direction.. 107

7.8 Required values of Fp……………………………………………………………. 111

7.9 Balance factor C………………………………………………………………….. 112

7.10 Relationship between RFW and Tcap……………………………………………. 120

7.11 RFW example for culvert ADA-41-0096………………………………………... 120

7.12 Comparison of controlling rating factors for design, current and worst case
conditions………………………………………………………………………… 124

8.1 Proposed load rating flow chart…………………………………………………... 127

8.2 Truncated pyramid showing how a surface load spreads over a base area at cover
depth……………………………………………………………………………… 130

8.3 A parabolic arch with hinged and fixed supports subjected to a uniform load…... 131

viii
8.4 Relationship between thrust due to dead or live load versus cover depth………... 157

8.5 Comparison of ODOT’s RFW to proposed RF for the worst possible culvert
condition………………………………………………………………………….. 163

8.6 Soil block mobilized by arching over culverts (FHWA HI-98-032, 2001)……… 165

8.7 Comparison of LRFD RFs to proposed RFs for the worst possible culvert
condition………………………………………………………………………….. 168

ix
x
CHAPTER 1

INTRODUCTION

1.1 Introduction and problem statement


ODOT maintains a large population of Corrugated Metal Pipe (CMP) and metal
arch culverts under highways. Each culvert has a different backfill height and is subjected
to different loading conditions. In order to evaluate the structural capacities of these
culverts, ODOT currently uses an in-house spreadsheet, CMP-Excel. This study was
initiated to verify and improve this program, ODOT’s current load rating procedures and
other in-house analytical tools. As a first step to achieve this objective, 39 in-service
culverts were tested under static and dynamic truck loading. Experimental data and
previous theoretical studies were used to evaluate and improve the current analysis
procedures.
Corrugated metal culverts have been used as drainage structures for many years,
especially in recent decades as short span bridge substitutes, for spans less than 50 ft
(15.24 m). The reasons for using corrugated metal culverts include: 1) light weight, 2)
economical, 3) generally maintenance free, 4) support heavy loads, and 5) relatively easy
handling and installation procedures. Generally large-span flexible metal culverts, which
have spans ranging between 10 to 50 ft (3.05 to 15.24 m), are constructed by bolting
together curved, corrugated metal plates. These culverts can have various shapes and
sizes as shown in Figure 1.1. One of the most popular corrugated metal culvert types has
been the pipe-arch culvert. The pipe-arch culverts do not require concrete footings and
are advantageous compared with other corrugated metal pipes in which head room is
limited. While these structures have generally performed well in the past, the reliability
of their field performances is uncertain.

1
Figure 1.1: Typical large-span culvert shapes.

One of the major factors affecting the performance of CMP culverts is corrosion.
The culvert structures are typically subjected to water inside and soil outside. Due to
frequent wetting and drying cycles, the culverts experience corrosion, especially near the
bottom surface. Similarly, as a result of mainly leakage of water along the bolted ends of
the plates on the sides and top of the culvert, the bolts and corrugated plates start
corroding. As the corrosion progresses, the strength and deformation capacity of the
culvert can be compromised. The corrosion damage inside the culvert can effectively be
minimized by casting concrete at the bottom or by using other available effective
rehabilitation techniques.
Other factors influencing with the performance of culverts include improper
placement of backfill material during construction, improper bedding, localized bulging
or buckling, excessive permanent deformations, and erosion. Although these problems do
not occur as common as corrosion in culverts, some culverts can fail as a result of these
factors which typically lead to structural distress in the culvert.
There have been many investigations dealing with the analysis and design of
corrugated metal culverts during the last four decades (Duncan, 1979; Chang et al., 1980;

2
Duncan et al. 1985; Alan, 1990; Manko and Beben, 2005). Laboratory testing, field
testing and analytical methods were utilized. Most of these studies were on performance
investigation of culverts under various backfill and external live load applications.
Research investigations on the response of existing culverts under static and dynamic
loads are very limited. The focus of most of the existing research is on field tests during
construction and under external static loads. As a result, it is hard to apply the published
field test results to design of new culverts and evaluation of existing culverts. In order to
investigate the performance of existing culverts under both static and dynamic loads, a
series of in-service culvert tests is needed to better understand the behavior of culverts
located in a similar small geographic area and designed using the same or similar
methods. For example, large number of identical culverts with different backfill heights
needs to be tested to investigate the effect of backfill height on the response of culverts
under static and dynamic loads. Similarly, the effect of other parameters, including the
culvert condition, shape and dimensions, can be investigated if all culverts are subjected
to tested comparable loads.
The American Association of State Highway and Transportation Officials
(AASHTO, 2002) provides general guidelines for structural design of metal culverts.
State Departments of Transportation typically develop additional design requirements or
adapt other supplementary guidelines such as those of the National Corrugated Steel Pipe
Association (NCSPA, 1995) for design of corrugated steel plate structures. According to
current design methods, culverts are generally designed based on a load rating factor,
which is the ratio of actual live load capacity to the required live load capacity. The
design procedure includes the load rating factor based on wall strength (RFW) and the
load rating factor based on minimum cover (RFC). The lower of these two factors
controls the culvert design. RFW is calculated from culvert wall thrust force resulting
from dead and live loads. However, moments created in the culvert plates are not
considered. The design procedures are based on the assumption that wall thrusts are
proportional to deflections. RFC is determined from the backfill height over the culvert.
However, the material properties of the culvert are not considered in the load rating. RFC
is independent from the thickness of corrugated metal plate as well as internal or external
loads. Consequently, current load rating method needs to be revised to develop a more

3
accurate design method, which can consider both RFW and RFC at the same time. To
ensure successful performance over the design life, corrugated metal culverts must be
designed and evaluated using an improved load rating method. The new method should
be based on theoretical approaches, and should be verified using the results from
experimental and numerical investigations.

1.2 Objectives
The overall objective of this study was to develop a new load rating method, and
to suggest recommendations for improving current load rating methods for corrugated
metal culverts. This objective was achieved by field-testing of various types of culverts,
computer simulation of several existing culverts, and an in-depth review of current design
practice and load rating procedures. An improved understanding of load rating method
would lead to a more accurate load rating design methodology for large-span corrugated
metal culverts. The main goals of this research were to:
• Review ODOT’s culvert database and previous field experiments conducted by
ODOT as well as other available sources to decide on a test matrix for testing and
analysis of a wide range of CMP and metal arch culverts in Ohio.
• Conduct analyses of typical culverts by taking into consideration all factors
involved, such as geometry, loading, backfill depth, soil-structure interaction, and
corrugated steel behavior. These analyses were carried out using standard ODOT
methods with CMP-Excel. The culverts were also analyzed using available theoretical
methods (e.g., ring compression theory), simplified procedures (e.g., AASHTO
Specifications), and interaction analysis programs (e.g., CANDE). These analyses helped
identify critical test parameters affecting the culvert response.
• Perform field testing of metal pipe and arch structures with different backfill
heights, damage and loading conditions, age, and size. A test matrix was designed to
investigate the effect of certain parameters such as backfill height on the response.
• Compare the field test results with solutions obtained from numerical and
analytical methods and tools used by ODOT, and determine for each loading condition
and back fill height if the experimental results were expected. To achieve this objective,

4
the calculated structural responses, including displacements, strains, moments and thrust
forces, were compared with those obtained from field testing.
• Modify/improve the current load rating programs or develop new methods, if
necessary, to analyze CMP and metal arch culverts. Develop recommendations based on
evaluation of observed and predicted responses by providing new analysis methods or
refinements of the existing programs and tools used by ODOT.

1.3 Scope and organization


In order to evaluate and improve the current load rating methods for metal
culverts, a comprehensive experimental research including field testing of 39 culverts
was conducted. First, the details of the experimental research are presented in this report.
Then, the test results are compared with the response predicted from numerical
simulations and theoretical methods. Current design practices and load rating procedures
are evaluated using the test data. Finally, recommendations are provided to improve the
current load rating procedures.
Chapter 2 presents a detailed review of previous flexible corrugated metal culvert
experiments, theoretical studies, analytical approaches, and design methods. Chapter 3
presents the details of the test setup, loading and instrumentation used in the experimental
tests and the description of 39 test culverts. Chapter 4 presents field test results including
culvert deflections and transverse and longitudinal strains measured by displacement
sensors and strain gauges, respectively. The experimental behavior of culverts under
static and dynamic loads is discussed in this chapter. Detailed data measurements from
the test culverts are provided in Appendices A and B. Evaluation of the 39 field tests and
the implication of test results are presented in Chapter 5. The effects of parameters,
including backfill height, truck loading, culvert condition and size, on the culvert
behavior are also discussed in this chapter.
Chapter 6 presents two-dimensional finite element analysis results from the
computer program CANDE. Some of the test culverts are modeled by using full mesh.
Predicted structural responses for the culverts are compared with those obtained form
experimental tests. The culvert design methods and load rating procedures currently used
by ODOT are evaluated in Chapter 7. Many discussions are conducted regarding current

5
ODOT’s load rating procedures. Chapter 8 presents a proposed load rating method to
improve current load rating procedures. Lastly, recommendations for an improved design
procedure, and summary and conclusions are presented in Chapter 9.

6
CHAPTER 2

BACKGROUND INFORMATION AND PREVIOUS RESEARCH

This chapter reviews published research that is relevant to the proposed work.
Most of the presented information is concerned with experimental, numerical and
theoretical work associated with behavior or design of culverts. However, little work has
been published on the load rating evaluation of corrugated culverts. Many years of field
experience have resulted in simplified theories and empirical formulae for the design of
buried flexible structures. These widely used classical methods were generally confirmed
by field observations and performance of culverts. The summary of considerable amount
of research on the behavior of various flexible metal culverts is outlined. Finally, more
rational modern methods, which have been applied recently to analyze flexible culverts,
are discussed.

2.1 Experimental studies


The behavior of full size culverts has been investigated in a number of field
studies over the past four decades. Most of these investigations were concerned with
circular, elliptical or arched long-span culverts. Most of the researchers attempted to
gauge culvert performance by measuring deformation of the culvert, and stress in the
culvert plate due to bending moments and axial thrusts. These investigations reported
culvert responses both during and after the construction.

2.1.1 Culvert deflections


The simplest method of measuring culvert deformation is to use surveying
techniques to monitor movements of control points marked on the culvert plate. Vertical
control is maintained with a surveyor’s level while a steel tape is used to measure chord
lengths between the control points. With triangulation, the changing coordinates of the

7
control points can be deduced (Bacher and Kirkland, 1986; Seed and Ou, 1986; Selig and
Musser, 1985; Kay and Flint, 1982). Unfortunately, this technique is not precise. Other
researchers have used dial gauges mounted on reference frames to track culvert
deflections (Gorman, 1981). Generally, such data can not be decomposed into deflections
normal and tangential to the culvert plate. More precise measurements of culvert
deformation can be accomplished using electrical transducers. Currently, such
transducers or displacement sensors are widely available. This scheme uses cables that
are stretched across various culvert diameters (Webb et al., 1999; Hurd and Sargand,
1988; Seed and OU, 1986; Bakht, 1985; Beal, 1982; Duncan and Jeyapalan 1982; McVay
and Selig, 1982) or attached to a reference frame inside the culvert (Bacher and Kirkland,
1986; Selig, 1975; Selig and Calabrese, 1975).

2.1.2 Corrugated plate forces


Bending moments and axial thrusts in the culvert plate can be determined from
strain measurements. Some researchers used weldable type strain gauges (Gorman, 1981;
Selig et al., 1979) while most current researchers use bonded electrical resistance strain
gauges. Unfortunately, several investigators (Bakht, 1981, 1985; Selig and Musser, 1985)
overlooked the influence of longitudinal strain in the culvert plate. Because of the
bellows action of the corrugated section, small longitudinal bending moments may cause
significant, localized strains transverse to the circumferential direction. Due to the
Poisson effect, longitudinal strains must be considered when trying to deduce the
circumferential stress resulting from bending moments and thrust forces.

2.1.3 Buckling
A characteristic feature of metal culverts is bending flexibility, which may cause
buckling. In an early study on long-span metal culverts, Meyerhof and Baikie (1963)
investigated the effect of soil on the strength of corrugated metal culverts. They reported
that this buckling failure of flexible plates depends on several critical parameters such as
coefficient of soil reaction or modulus of deformation of the soil, and the flexural rigidity
of the plates. The observed ultimate loads and modes of failure of the plates were in

8
reasonable agreement with the estimates and also support the ring compression theory,
which will be described later in this chapter.
Buckling of culvert structures was also studied by Ghobrial and Adbel-Sayed
(1985). In their experimental study, they considered the formation of plastic hinges in the
culvert walls. They observed a sudden snap-through failure for large span and shallow
backfill height. Short span culverts with deep backfill soil did not exhibit this response,
but experienced increasing displacements after each load increment. In addition, the first
plastic hinge formed at the crown. The formation of the second plastic hinge was
observed farther away from the crown.

2.2 Analytical studies


Analytical or numerical solution techniques have been used to predict the load
distribution around the culvert as well as the overall behavior of culvert. In order to
obtain an exact solution, many culverts have been investigated for analytical approaches
by numerous researchers (Savin, 1961; Burns, 1965; Dar and Bates, 1974; Bakht and
Agarwal, 1988). Most of these studies aimed at calculating the stress state around a hole
inside an elastic solid block using several simplifying assumptions. Besides these
traditional methods, this report presents and uses numerical methods based on more
powerful computational techniques, such as finite difference or finite element methods.
Because the restrictions and approximations for applying the solutions obtained from
elastic theory are no longer required when more robust numerical techniques that can
incorporate nonlinear material behavior are available. The numerical approaches to
analyze soil-culvert structures may be distinguished with two approaches. The first
represents the soil system as a set of springs to simulate the soil response. The second
approach deals with the soil medium as a continuum, and the soil-culvert system is
analyzed by the finite element method.

2.3 Theoretical studies


In order to capture the behavior of culverts accurately, soil-structure interaction
has to be considered during the design process. However, in traditional methods of design,

9
usually the soil structure interaction is not properly represented. In this section, the
theoretical studies for design are briefly summarized.
In general, design of most flexible culverts is based on three criteria: a deflection
criterion, a wall compression criterion, and a wall buckling criterion. The deflection
criterion is based on a deflection originated by the Iowa deflection formula suggested by
Spangler (1941), and modified by Masada (2000) to get rid of many assumptions used to
calculate vertical deflection. The wall and seam compression criterion depends on a
design compression force obtained from the ring compression theory as proposed by
White and Lager (1960) and adopted with some empirical modifications by a few codes
like AISI handbook (1994). The wall buckling criterion follows the studies conducted by
many researchers, such as Watkins and Anderson (1999), Moore et al. (1988), Meyerhoff
(1963), and Abdel-Sayed (1978).

2.3.1 Soil-structure interaction studies


The pioneering work regarding soil-structure interaction was developed by
Marston (1930) and Spangler (1941, and 1973). The Iowa formula, described in Section
2.3.2, was based on their experimental and theoretical studies for calculating pipe
deflection under earth load. Their studies had focused on the interaction between flexible
tubular structures and the surrounding soil. Their work was a trigger to develop soil-
structure interaction approaches.

2.3.2 Marston-Spangler method


The Marston-Spangler approach (Marston, 1930; Spangler, 1941 and 1973) was
originally developed for round culverts with small diameter. It is based on the assumption
that vertical soil pressures at the top and bottom of the pipe are uniform, and the sides of
the pipe are subjected to parabolic horizontal pressure. These soil pressures are illustrated
in Figure 2.1. As shown in this figure, the vertical pressure is uniformly distributed over
the pipe diameter and the horizontal pressure over the arc length corresponding to a 100°
angle at the center.
The intensity of the effective uniform pressure is assumed to be equal to the
weight of a vertical soil column above the crown of the culvert plus the pressure applied

10
at the surface. The intensity of horizontal soil pressure on the pipe is assumed to be
proportional to the horizontal movement of the pipe. The maximum horizontal pressure
on the sides is up to 35 percent greater than the vertical pressure applied at the top as
shown in Figure 2.1.

Figure 2.1: Pressure distribution assumed in the Marson-Spangler method.

According to the assumed pressure distribution, the circumferential thrust is a


function of the radius of the circular cross-section of the culvert, R , and the intensity of
the vertical pressure at the crown, Pc . The thrust varies from a minimum of a 0.7 Pc R at

the top and bottom of the culvert and Pc R at the sides, to a maximum of about 1.1 Pc R at
the haunches. The corresponding circumferential moment in the culvert wall varies from
about 0.02 Pc R 2 at the top, sides and bottom, to about -0.02 Pc R 2 at the haunches
(Meyerhoff, 1968; Spangler, 1941).
The radial deflection under the assumed loading consists of a downward
movement of the top and outward movements of the sides of the culvert. The maximum
horizontal deflection, d in inches, is given by:
kl Pc R 4
d= (2.1)
EI + 0.061E ' R 3
where E denotes modulus of elasticity of the culvert wall material (psi), I denotes
moment of inertia per unit length of cross section of culvert (in.4/in.), E ' denotes

11
modulus of soil reaction, and kl is a constant that depends on the bedding angle and a

deflection lag factor.


Spangler (1941) concluded from his experimental investigation that, if the vertical
diameter of a circular flexible culvert decreases by about 20% from the initial diameter,
the pipe is in a state of incipient collapse. Additional vertical load on the pipe causes
failure due to reversal curvature or snap-through buckling. It has become customary,
since then, to refer to failure condition in flexible pipe as 20% decrease in vertical
diameter. It was also observed that the decrease in vertical diameter is almost equal to the
increase in the horizontal diameter and can be obtained from Equation 2.1. The final
design is achieved by adjusting the in-plane bending stiffness of the culvert to limit the
vertical deflection to five percent or less of the diameter.

2.3.3 Ring compression theory


White and Layer (1960) suggested that a flexible culvert can be analyzed as a thin
ring in compression when the flexible culvert is installed in a compacted backfill. The
theory was developed based on the assumption that the non-uniform soil pressure
distribution around the culvert has little effect on the magnitude and distribution of the
circumferential thrust. Thus, it simplified the complex loading condition by neglecting
the effect of any soil friction forces and assuming a uniform pressure distribution. This
assumption is considered valid for flexible circular culverts with cover heights exceeding
one-eighth of the culvert diameter. The uniform pressure Pc is taken as the over-burden

pressure plus any distributed live load, Pl including impact at the top of the culvert:

Pc = γ h '+ Pl (2.2)

where γ denotes unit weight of backfill material, h ' denotes average backfill height
above the structure, and Pl denotes the equivalent live load pressure. For backfill heights

exceeding the diameter of the culvert, the minimum backfill height, h at the top of the
structure may be used instead of the average height h ' in Equation 2.2. The average and
minimum cover heights are shown in Figure 2.2.

12
Figure 2.2: Pressure distribution assumed by the ring compression theory for a pipe
culvert.

According to the ring compression theory, the circumferential thrust in the culvert
wall is given by:
⎛S⎞
T = Pc ⎜ ⎟ (2.3)
⎝2⎠
where S denotes span of the culvert cross-section, which is equal to 2R for circular
pipes.
In the wall of a non-circular culvert, the radial soil pressure is considered to vary
in such a way that the circumferential thrust in the wall remains constant throughout the
circumference. Hence, the soil pressure on the structure at any point may be written as:
T
P= (2.4)
R'
where, R ' denotes radius of curvature at the point under consideration.
The soil pressure, being inversely proportional to the radius of curvature, is
maximum at the point of minimum radius, for example, at the haunches of a pipe arch.
The distribution of soil pressure according to the ring compression theory for a pipe arch
culvert is shown in Figure 2.3. For any culvert shape, a successful design is achieved by

13
supplying sufficient culvert wall area such that the thrust stress is smaller than the wall
strength.

Figure 2.3: Pressure distribution assumed by the ring compression theory for a pipe arch
culvert.

2.3.4 Buckling considerations


Buckling can be defined as the loss of resistance of culvert wall resulting from
excessive flexural deformation. Generally, design against buckling is achieved by
selecting an appropriate wall thickness so that it is capable of carrying the thrust force
with an adequate factor of safety against buckling stress. Booy (1957) reported first that
the culvert wall may fail by buckling before culvert deflects enough to collapse, and
proposed the buckling formula derived for an elastically supported beam (Timoshenko,
1964) which results in the following equation:

2 kEI
fb = (2.5)
A 1 −ν 2
where fb denotes buckling stress, A denotes cross-sectional area per unit length of the
culvert wall, E denotes modulus of elasticity of plate, I denotes moment of inertia of
plate, k denotes coefficient of soil reaction, and ν denotes the Poisson’s ratio of the
plate. The buckling strength of soil-culvert structures is significantly improved by the soil
support which provides resistance to both inward as well as outward movement (Moore

14
et al., 1988). Buckling theory has been reviewed by Watkins (1960), Abdel-Sayed (1978),
Baikie and Meyerhof (1982), Moore (1989), and Haggag (1989). According to their
research, theoretical approaches to solve the buckling problem can be classified into two
methods; one is linear theory, and the other is single-wave theory. The linear theory is to
solve the buckling problem by considering the linear buckling strength as buckling occurs
around the culvert circumference. The single-wave theory takes into account only one
local buckling in the culvert wall. Moore (1987) investigated both theories and concluded
that the linear theory is more appropriate for culverts.

2.3.5 Live loads


Culverts are frequently installed under highway and airport pavements and
railroad tracks. If a live load is close to the culvert, the load is added to the other loads
which the culvert must carry. However, due to the distributional effect of backfill height
over the culvert, the effect of live loads may be negligible when culverts are installed
more than eight feet deep.
Existing theory and design procedures for load distribution through soil medium
are based upon the theories of Boussinesq (1883) and AASHTO (2002). Theoretical
solutions for the load distribution are based on the theories of elasticity and specific
assumptions relative to material properties and boundary conditions. The theoretical
bases are reviewed in the following section.

2.3.5.1 Boussinesq equations


Boussinesq (1883) developed equations for the problem of stress due to a point
load applied to an infinitely large plane with a homogenous, elastic, and isotropic
material. The equations were developed for the stress under a point load applied to the
surface of an infinite wide plane as shown in Figure 2.4. The equation for vertical stress
is:
3P z 3
Δσ z = (2.6)
2π L5

where L = x 2 + y 2 + z 2 , and P is a point load.

15
L

Figure 2.4: Vertical stress caused by a point load.

2.3.5.2 Load distribution assumptions


Pavements designed for heavy truck traffic substantially reduce the pressure
transmitted through a wheel to the underlying culvert. As discussed previously,
Boussinesq equation can be used to estimate the pressure intensity. A more practical
method as accepted by AASHTO was developed. The maximum highway wheel loads
generally considered for design purposes are those specified by AASHTO for HS-20
truck and alternate load configurations. For HS-20 wheel load, 16000 pounds are carried
on dual wheels, and the contact pressure is 80 pounds per square inch. Therefore, the
contact area is 200 square inches with a length 20 inches and a width of 10 inches.
According to AASHTO, at depths, H , below the surface, the wheel pressure is assumed
to increase by the value of 1.75H as shown in Figure 2.5.

16
Figure 2.5: Distributed load area for single dual wheel of HS-20 truck.

17
CHAPTER 3

EXPERIMENTAL RESEARCH AND CULVERT INFORMATION

3.1 Introduction
In order to better understand the behavior of existing corrugated metal culverts,
39 culverts were tested using a heavily loaded truck that represents a load larger than
typical design live load. In the selection of test culverts, several factors including the
backfill height, current condition or damage, age, span length, and culvert geometry were
considered.
This chapter describes the field test setup, instrumentation, and test culverts.
Culvert deflections and strains were measured by using electric displacement sensors and
strain gauges. Deflections and strains of each culvert wall were recorded using a mobile
data acquisition system connected to a laptop computer. The data acquisition system had
20 channels, which can receive data from an array of displacement sensors and strain
gauges installed on the culvert. The frequency of measurement was ten data points per
second, which provided sufficient accuracy during the dynamic tests. All electrical cables
were terminated with connectors that allowed them to be easily connected and
disconnected to the data acquisition system to facilitate set up in the field. This chapter
also includes general description and unique properties of 39 existing culverts tested. The
culverts were located in 24 different counties in Ohio.

3.2 Instrumentation

3.2.1 Deflections
The most direct method to investigate culvert deformations is to measure
deflections (i.e., displacements) at several control points on the culvert plates. Five HS-
100 displacement transducers and one HS-50 displacement transducer manufactured by

18
the Vishay Measurement Company were used to measure culvert deflections and slippage
between the metal plates in this project. The sensors had a maximum displacement
capacity of ±50 mm (HS-50 for slippage) or ±100 mm (HS-100 for deflection).
Two portable aluminum instrumentation frames were constructed and the frame
placed inside the test culvert few days before the experiment. Usually, two culverts were
tested during the same day. The displacement sensors were installed on the
instrumentation frame and calibrated prior to testing. Figure 3.1 shows a photograph of
the instrumentation frame, which consists of five arms with a displacement sensor
mounted on each (channels CH#1-5). The body of the frame was about 1.5 m long and
0.9 m tall. The arms were constructed using rectangular hollow aluminum sections that
can slide inside each other and can thus be extended from 1.2 to 4.6 m, depending on the
height of the culvert. Using clamps and bolts, each arm can be attached to the body at the
desired angle within a range of ±10 degrees from the original position. The modular
construction of the frame allows for rapid assembly and disassembly in the field. The
sixth sensor was installed between the ends of two plates connected at the crown of the
culvert to measure potential slippage between the plates during the tests (CH #6). No
joint slip was recorded at the crown during the tests.

3.2.2 Strains
A total of 14 channels were used to monitor changes in strains on the culvert
walls in the transverse (nine gauges) and longitudinal (five gauges) directions (Figure
3.2). Type CEA-06-250UW-120 electrical resistance uni-directional strain gauges
manufactured by Vishay were used. All gauges had a nominal resistance of 120 ohms and
a nominal gauge factor of 2.095 with a nominal transverse sensitivity of 0.3 percent.
Twelve gauges were placed on the same transverse section of the culvert directly under
the center of gravity of the test truck. The other two gauges were installed 1 ft (0.3 m)
and 2 ft (0.6 m) away from the center of gravity of the test truck to monitor the strain
variations in the longitudinal direction.

19
(a)

(b)

Figure 3.1: a) instrumentation frame, and b) displacement sensor locations.

20
Gauges were positioned with axes oriented parallel and perpendicular to the
direction of the plate corrugation to measure strains in the circumferential and
longitudinal directions, respectively. In order to obtain more reliable strain measurements,
six strain gauges were mounted near the crown of the culvert. The gauges CH #07, CH
#10, and CH #11 were installed inside the “peak” of the corrugation and gauges CH #09
and CH #12 were installed on the “valley” of the corrugation (Figure 3.2). Gauge CH #08
was installed at the middle of the corrugation.

3.3 Live load application


Most culvert structures were designed for the AASHTO HS20-44 truck with a
total weight of 72 kips. In this research, culverts were tested using trucks representative
of either the HS20-44 truck or ODOT Buried Structures Research Truck, BUSRET with a
total weight of 63 kips. The specified axle loads for the HS20-44 and ODOT BUSRET
trucks are provided in Table 3.1. The trucks were loaded typically with crushed stone at a
local ODOT facility and then driven to the testing sites.
Each culvert was tested twice (once in each direction using the same driving lane)
for ten different static load cases (i.e., positions) and six different dynamic load speeds.
The test truck was positioned at ten locations over the culvert to produce the static loads
LC 1 through 10 as shown in Figure 3.3. The truck was then driven over the culvert at
speeds ranging from 5 to 40 mph, or the maximum legal speed, to investigate the
response to dynamic loading.

21
Figure 3.2: Strain gauge locations (circled channels indicate a strain gauge placed in
longitudinal direction of culvert).

Table 3.1: HS20-44 and ODOT BUSRET truck loads.

Truck weight (kips)


Truck
Steering axle Drop axle 1st drive axle 2nd drive axle Total weight
type
(kips) (kN) (kips) (kN) (kips) (kN) (kips) (kN) (kips) (kN)
HS20-44 8 36 - - 32 142 32 142 72 320
ODOT
12 52 17 76 17 76 17 76 63
BUSRET

22
Figure 3.3: Test truck positions for ten static load cases.

23
3.4 Description of test culverts
The 39 in-service corrugated metal culverts were investigated in this study, the
locations of which are shown in Figure 3.4. Table 3.2 summarizes the geometric and
material properties of each culvert. Each culvert name includes ten characters: the first
three letters are the abbreviation of the county name, the second three numbers denote the
highway or state route number, and the final four numbers denote the straight line mile
(SLM) distance from the west or south border of the county to the culvert location. For
example, ADA-041-0096 represents a culvert located in Adams County, on State Route
41, and 0.96 miles from the west county border. The typical culverts in Ohio can be
classified as pipe-arch culverts, low-profile arch culverts, arch culverts, and pipe culverts
(Figure 1.1). The majority of culverts tested in this study were pipe-arch culverts,
whereas CLI-068-0942 and KNO-062-1480 were pipe culverts; DEL-037-0333 was a
low-profile arch culvert; and LAK-608-0303, and NOB-83-0419 were pipe culverts. All
culverts were constructed of 6 × 2 galvanized corrugated steel, which has a corrugation
pattern defined by a pitch of 6 in. (152.4 mm) and a depth of 2 in. (50.8 mm) as defined
by AISI (1994). Differences in cross sectional area and moment of inertia (Table 3.2)
arise from different thicknesses of the corrugated metal plates. The Ohio Department of
Transportation (ODOT) annually inspects each culvert in Ohio with a span exceeding 10
ft (3 m) and assigns a general appraisal number N to reflect its current condition. An
appraisal number equal to 1 represents the worst condition and a value of 9 indicates the
best possible condition for a given culvert. Appraisal numbers provided in Table 3.2
correspond to the time of testing.
A brief description of 39 test culverts and their unique properties are reported
below. The photographs of the culverts were taken within a few days of the testing day.
The reported culvert damage is based on the observations of the researchers. Although
the observed damage was mostly in line with the appraisal number reported by ODOT, in
some cases the reported and observed damage were conflicting.

24
Figure 3.4: Corrugated metal culvert locations tested.

25
Table 3.2: Characteristics of corrugated metal culverts at 39 test sites.

Year Culvert Culvert Backfill Plate Section Moment of General


* * *
built Span Height height thickness area inertia appraisal
Culvert name
number
(ft) (ft) (ft) (in.) (in.2/ft) (in.4/ft) N
ADA-041-0096 1952 11.8 7.5 4.0 0.170 2.449 1.154 4
ALL-309-1664 1955 16.7 10.0 5.3 0.218 3.199 1.523 5
ALL-696-0410 1957 12.7 8.0 2.6 0.188 2.739 1.296 5
ATB-193-2336 1961 15.9 9.8 3.3 0.249 3.658 1.754 4
ATB-307-1801 1962 16.6 10.0 4.3 0.218 3.199 1.523 4
CLI-068-0942 1938 11.5 11.0 22.3 0.249 3.658 1.754 7
COL-170-1981 1957 13.8 7.5 0.9 0.218 3.199 1.523 5
COL-172-1209 1966 16.7 10.0 2.6 0.218 3.199 1.523 5
DEF-018-1560 1981 10.5 6.9 2.8 0.111 1.556 0.725 5
DEL-037-0333 1976 21.0 6.8 16.0 0.188 2.739 1.296 5
FAY-734-0790 1958 12.5 8.0 1.8 0.188 2.739 1.296 5
FRA-062-2688 1958 12.5 7.8 5.3 0.170 2.449 1.154 5
GAL-218-0578 1956 15.7 9.1 5.0 0.249 3.658 1.754 5
HIG-134-0782 1958 12.7 7.8 2.1 0.188 2.739 1.296 5
HOC-093-1975 1960 14.8 8.5 4.3 0.188 2.739 1.296 3
HOC-664-1172 1963 14.6 9.0 2.5 0.188 2.739 1.296 5
JAC-327-0077 1956 13.5 8.5 2.5 0.218 3.199 1.523 5
KNO-013-2162 1963 14.3 8.5 7.4 0.188 2.739 1.296 6
KNO-062-1480 1960 12.0 12.5 24.5 0.218 3.199 1.523 6
LAK-608-0303 1961 14.2 7.3 2.0 0.218 3.199 1.523 4
LIC-079-2276 1981 10.3 6.8 3.3 0.111 1.556 0.725 6
LIC-661-0805 1982 20.0 12.5 5.7 0.170 2.449 1.154 6
MAD-038-0935 1961 10.7 6.9 1.9 0.170 2.449 1.154 3
MAD-056-1490 1956 14.7 8.8 6.8 0.188 2.739 1.296 6
MAD-665-1071 1964 12.5 8.0 3.3 0.170 2.449 1.154 6
MAR-229-0093 1972 16.6 10.0 3.0 0.218 3.199 1.523 5
MAR-309-1248 1968 10.7 6.8 2.5 0.170 2.449 1.154 7
MAR-423-0214 1972 12.5 7.9 5.5 0.170 2.449 1.154 6
MAR-423-0379 1972 16.8 9.8 3.8 0.170 2.449 1.154 7
MRW-019-0456 1997 12.8 8.0 3.0 0.140 2.003 0.938 9
MRW-095-1637 1966 11.0 6.9 2.0 0.170 2.449 1.154 7
MRW-095-1796 1970 12.5 7.9 8.2 0.170 2.449 1.154 7
NOB-083-0419 1941 17.0 7.6 1.5 0.188 2.739 1.296 4
PIC-138-0091 1981 13.4 7.6 3.0 0.111 1.556 0.725 2
PIC-159-0868 1989 12.0 12.0 8.9 0.188 2.739 1.296 8
PIC-752-0748 1981 18.0 11.5 9.0 0.170 2.449 1.154 8
SAN-412-0575 1933 13.0 7.0 1.5 0.170** 2.449 1.154 5
SAN-510-0412 1960 11.5 7.4 1.4 0.218 3.199 1.523 5
TRU-193-0215 1954 14.2 9.3 4.6 0.218** 3.199 1.523 3
* The size of culvert and backfill height is measured during inspection.
** These values are assumed from the thickness measured at the field.

26
Table 3.2 continued

Year Culvert Culvert Backfill Plate Section Moment of General


* * *
built Span Height height thickness area inertia appraisal
Culvert name
number
(m) (m) (m) (mm) (cm2/m) (cm4/m) N
ADA-041-0096 1952 3.6 2.3 1.2 4.318 51.8 157.6 4
ALL-309-1664 1955 5.1 3.0 1.6 5.537 67.7 208.0 5
ALL-696-0410 1957 3.9 2.4 0.8 4.775 58.0 177.0 5
ATB-193-2336 1961 4.8 3.0 1.0 6.325 77.4 239.5 4
ATB-307-1801 1962 5.1 3.0 1.3 5.537 67.7 208.0 4
CLI-068-0942 1938 3.5 3.4 6.8 6.325 77.4 239.5 7
COL-170-1981 1957 4.2 2.3 0.3 5.537 67.7 208.0 5
COL-172-1209 1966 5.1 3.0 0.8 5.537 67.7 208.0 5
DEF-018-1560 1981 3.2 2.1 0.9 2.819 32.9 99.0 5
DEL-037-0333 1976 6.4 2.1 4.9 4.775 58.0 177.0 5
FAY-734-0790 1958 3.8 2.4 0.5 4.775 58.0 177.0 5
FRA-062-2688 1958 3.8 2.4 1.6 4.318 51.8 157.6 5
GAL-218-0578 1956 4.8 2.8 1.5 6.325 77.4 239.5 5
HIG-134-0782 1958 3.9 2.4 0.6 4.775 58.0 177.0 5
HOC-093-1975 1960 4.5 2.6 1.3 4.775 58.0 177.0 3
HOC-664-1172 1963 4.5 2.7 0.8 4.775 58.0 177.0 5
JAC-327-0077 1956 4.1 2.6 0.8 5.537 67.7 208.0 5
KNO-013-2162 1963 4.4 2.6 2.3 4.775 58.0 177.0 6
KNO-062-1480 1960 3.7 3.8 7.5 5.537 67.7 208.0 6
LAK-608-0303 1961 4.3 2.2 0.6 5.537 67.7 208.0 4
LIC-079-2276 1981 3.1 2.1 1.0 2.819 32.9 99.0 6
LIC-661-0805 1982 6.1 3.8 1.7 4.318 51.8 157.6 6
MAD-038-0935 1961 3.3 2.1 0.6 4.318 51.8 157.6 3
MAD-056-1490 1956 4.5 2.7 2.1 4.775 58.0 177.0 6
MAD-665-1071 1964 3.8 2.4 1.0 4.318 51.8 157.6 6
MAR-229-0093 1972 5.1 3.0 0.9 5.537 67.7 208.0 5
MAR-309-1248 1968 3.3 2.1 0.8 4.318 51.8 157.6 7
MAR-423-0214 1972 3.8 2.4 1.7 4.318 51.8 157.6 6
MAR-423-0379 1972 5.1 3.0 1.2 4.318 51.8 157.6 7
MRW-019-0456 1997 3.9 2.4 0.9 3.556 42.4 128.1 9
MRW-095-1637 1966 3.4 2.1 0.6 4.318 51.8 157.6 7
MRW-095-1796 1970 3.8 2.4 2.5 4.318 51.8 157.6 7
NOB-083-0419 1941 5.2 2.3 0.5 4.775 58.0 177.0 4
PIC-138-0091 1981 4.1 2.3 0.9 2.819 32.9 99.0 2
PIC-159-0868 1989 3.7 3.7 2.7 4.775 58.0 177.0 8
PIC-752-0748 1981 5.5 3.5 2.7 4.318 51.8 157.6 8
SAN-412-0575 1933 4.0 2.1 0.5 4.318** 51.8 157.6 5
SAN-510-0412 1960 3.5 2.3 0.4 5.537 67.7 208.0 5
TRU-193-0215 1954 4.3 2.8 1.4 5.537** 67.7 208.0 3
* The size of culvert and backfill height is measured during inspection.
** These values are assumed from the thickness measured at the field.

27
ADA-041-0096
This pipe-arch culvert was installed in 1952 near Bradysville in Adams County, Ohio.
The roadway over the culvert has a vertical slope of approximately 5 degrees. The culvert
was fabricated using eight corrugated steel plates and is perpendicular to the centerline of
the road. Before the tests, the culvert was inspected visually. The culvert had significant
corrosion problems at several locations on the plates mostly concentrating around
connections and bolts. This culvert was tested on November 13, 2006. The condition of
the culvert is shown in Figure 3.5(a).

ALL-309-1664
The culvert was constructed in 1955, and was located south of Lafayette in Allen County,
Ohio. The pipe culvert was installed perpendicular to the roadway. The culvert was
fabricated using nine corrugated steel plates. Most of the bolts at the top plate were
corroded as shown in Figure 3.5(b). A large part of two plates at the bottom had
significant corrosion damage. Some bolts on these plates were damaged and lost. The
testing date of this culvert was September 17, 2007.

ALL-696-0410
The pipe-arch culvert installed in 1955 was located on State Route 696, south of Bluffton
in Allen County, Ohio. The culvert had a skew angle of 4 degrees from the perpendicular
line to the roadway. Eight corrugated plates were bolted together to fabricate the arch
shape. Figure 3.5(c) shows the damaged parts of the culvert. As shown in the figure, most
of the bolts at the crown are slightly corroded. Water leakage was observed around the
corroded bolts. Four plates at the bottom are also corroded widely. Some bolts on these
plates have been damaged and lost. However, overall appearance and structural integrity
of the culvert appeared to be fine. The culvert was tested with the culvert ALL-309-1664
on the same day.

ATB-193-2336
The corrugated steel culvert tested on August 16, 2007 was installed in 1961 and located
on State Route 193 near Kingsville in Ashtabula County, Ohio. The culvert was

28
perpendicular to the roadway and consisted of nine corrugated plates bolted to make the
shape of pipe-arch. There was extensive corrosion damage inside the culvert. Four
structural plates at the bottom of the culvert were corroded. Generally, corrosion in steel
culverts occurs at or near bolts. However, this culvert had corroded bolts as well as
extensive corrosion in steel plates around the corroded bolts. Figure 3.5(d) shows the
pictures of the damage in the culvert.

ATB-307-1801
This culvert was also tested the same day with the culvert ATB-193-2336. The
corrugated metal culvert was installed in 1962, on State Route 307, near the village of
Jefferson in Ashtabula County, Ohio. Skew angle of the culvert is 11 degrees. The shape
of the culvert was pipe-arch, and nine corrugated plates were bolted together. Extensive
steel corrosion was observed near the entrance of the culvert during our inspection.
However, the center of the culvert did not have significant corrosion damage except for
the bottom plates as shown in Figure 3.5(e).

29
(a) ADA-041-0096

(b) ALL-309-1664

(c) ALL-696-0410

(d) ATB-193-2336

(e) ATB-307-1801

Figure 3.5: The inspected condition of the culverts prior to testing.

30
CLI-068-0942
The corrugated metal round pipe was installed in 1938 and was tested on November 8,
2006. This was the oldest test culvert and was located on State Route 68, near the south
of Midland village in Clinton County, Ohio. The pipe culvert was perpendicular to the
roadway. There was a railroad right next to the roadway, and the culvert was connected
with a concrete tunnel, that is, half of the culvert consisted of metal one and the other part
was concrete one. The concrete tunnel was used to support heavy railway load and the
metal culvert was used for roadway. Although the behavior of concrete tunnel structure
may affect that of the culvert, it was assumed that they were not affecting each other. The
pipe culvert had eight corrugated plates to form circular shape. The bottom of the side
plates were supported by a concrete base. Some parts of two structural plates at both sides
of the culvert were corroded lightly as shown in Figure 3.6(a). Many bolts on the
corroded parts had damage due to steel corrosion. However, the culvert was in reasonably
good condition.

COL-170-1981
The installation year of this arch culvert was 1957. The culvert was located on State
Route 170, near the village of East Palestine in Columbiana County, Ohio. The skew
angle of the culvert was 28 degrees. Four corrugated plates were bolted together to
fabricate the arch shape (half circle) (see Figure 1.1). Concrete was cast near the bottom
to support the each end of the arch culvert. Some parts of metal plates near the concrete
supports were corroded. However, overall appearance of the culvert was very clean with
insignificant damage. The loading truck, portable data acquisition system, and culvert are
shown in Figure 3.6(b). The testing date of the culvert was August 27, 2007.

COL-172-1209
The corrugated steel culvert built in 1966 was located on State Route 172 near Lisbon in
Columbiana County, Ohio. The culvert was tested with the culvert COL-170-1981
together on the same day. The culvert was set perpendicular to the roadway, and consists
of nine corrugated plates to form the pipe-arch shape. Four structural plates at the bottom
of the culvert were mildly corroded as shown in Figure 3.6(c).

31
DEF-018-1560
The corrugated metal culvert was tested on September 12, 2007, and was located on State
Route 18 near Defiance, Ohio. It had the skew angle of 9 degrees from the perpendicular
line to the roadway. The culvert was fabricated using seven corrugated steel plates.
Figure 3.6(d) shows general view of the culvert. The bottom two plates were slightly
corroded. Overall appearance of the culvert was good.

DEL-037-0333
This culvert was the only low profile culvert in this study. It was tested on October 25,
2006. The culvert having a low, flattened arch appearance was constructed in 1973, on
State Route 37, near the north of the Warrensburg village in Delaware County, Ohio. The
culvert was installed perpendicular to the centerline of the road. The culvert was
inspected visually before the test. The culvert seemed to be installed on a big flat rock.
Some parts of the rock were broken inside the culvert. Few corroded bolts at the crown
were found. The shape of the culvert was not symmetrical at the time of testing because
one side has settled and deflected more than the other side as shown in Figure 3.6(e).

FAY-734-0790
The corrugated steel culvert was installed in 1958 on State Route 734 near Jeffersonville
in Fayette County, Ohio. This culvert was installed perpendicular to the centerline of the
road. Seven corrugated plates were bolted together to fabricate the pipe-arch shape.
Figure 3.6(f) shows general view of the culvert. Bottom structural plates of the culvert
were damaged due to steel corrosion. The connections between plates were detached and
created slight gaps at the entrance of the culvert. Water flow passing through the culvert
was fast. However, the overall appearance was good. This culvert was tested on June 9,
2006.

FRA-062-2688
The installation year of the corrugated pipe-arch culvert was 1958. The culvert, tested on
July 17, 2007 was located on State Route 62 near New Albany in Franklin County, Ohio.

32
The skew angle of the culvert was 47 degrees from the perpendicular line to the roadway.
In order to form the pipe-arch shape, eight corrugated metal plates were bolted together.
There were corroded plates at the bottom. However, the overall appearance seemed good.
Lots of sand and gravel deposits were piled up at one side of the culvert bottom as shown
in Figure 3.6(g). The sand and gravel appeared to be brought inside the culvert by the
water flow rather than leakage of the backfill material outside the culvert.

GAL-218-0578
The culvert located on State Route 218 near Mercerville, Ohio was built in 1956. The
culvert was installed with a skew angle of 33 degrees from the perpendicular line to the
roadway. There were deposits of sand and gravel inside the culvert. A localized bulge
was found at the crown of the culvert. Repair traces and steel corrosion around the bulge
were found. Some parts of the crown were severely corroded as shown in Figure 3.6(h).
This culvert was tested on September 27, 2007.

HIG-134-0782
Installed in 1985 was the corrugated steel culvert on State Route 134 near Buford in
Highland County, Ohio. The skew angle was 34 degrees from the perpendicular line to
the centerline of the road. Figure 3.6(i) shows the overall view as well as damaged parts
of the culvert. As it can be seen in the figure, concrete was cast at the bottom of the
culvert below the spring line to prevent damage and corrosion and to enhance the strength
of culvert by providing better support near the bottom of the culvert. The culvert had a lot
of damage concentrations at the plate connections due to corrosion. However, the
damaged parts were not under the roadway but under the soil deposit. Thus, the observed
damage has not affected probably the structural response of the culvert greatly.

33
(a) CLI-608-0942 (b) COL-170-1981 (c) COL-172-1209

(d) DEF-018-1560 (e) DEL-037-0333

(f) FAY-734-0790 (g) FRA-062-2688

(h) GAL-218-0578

(i) HIG-134-0782

Figure 3.6: The inspected condition of the culverts before testing.

34
HOC-093-1975
The corrugated metal culvert was installed in 1960, on State Route 93, near the village of
Logan in Hocking County, Ohio. The culvert had a skew angle of 26 degrees from the
perpendicular line to the roadway. Eight corrugated plates were bolted together to
fabricate the pipe-arch shape. There were extensive deposits of sand and gravel inside the
culvert as shown in Figure 3.7(a). Both corners at the bottom of the culvert were
damaged due to corrosion. However, the overall condition of the culvert seemed to be
good. The culvert was tested on July 9, 2007.

HOC-664-1172
The installation year of the culvert was in 1963 and tested on July 9, 2007. The skew
angle of the culvert was 26 degrees. In order to form the culvert, eight corrugated plates
were bolted together. A deposit of gravel in the culvert was found during culvert
inspection. Steel corrosion damage was also found at the bottom of the culvert, the plate
connections and at bolts on these connections. The overall condition of the culvert is
shown in Figure 3.7 (b).

JAC-327-0077
The culvert built in 1956 is on State Route 327 near Wellston in Jackson County, Ohio.
The skew angle was 15 degrees from the perpendicular line to the centerline of the road.
Concrete was cast at the bottom of the culvert as shown in Figure 3.7(c). There were
extensive deposits of sand and gravel inside the culvert. The deposits filled the bottom of
the culvert up to the spring line. It is possible that the sand and gravel might had been
dumped during or after construction because similar sand and gravel deposits could not
be found in the vicinity of the culvert. The culvert had widespread damage at the plate
connections due to corrosion. The testing date was September 27, 2007.

KNO-013-2162
The corrugated metal culvert with a skew angle of 9 degrees was constructed in 1963,
and tested on October 16, 2006. It was located on State Route 13, near Fredericktown in
Knox County, Ohio. Seven corrugated plates were bolted together to form the culvert.

35
The bottom of the culvert is seated into concrete as shown in Figure 3.7(d). Although the
general appraisal number was 6, the culvert looked very clean. No rust and damage were
found. The excellent condition of this 45 year old culvert shows that corrosion may not
be a problem for some culverts in some locations.

KNO-062-1480
The corrugated metal pipe was installed in 1960, and is located on State Route 62, near
the village of Danville. This was the deepest culvert tested in this study. The backfill
height of the culvert was 24.5 ft. The skew angle of the culvert is 18 degrees. The circular
pipe consists of six corrugated plates. At the bottom of the culvert, concrete was cast to
prevent steel corrosion as shown in Figure 3.7(e). Inside of the culvert was clean. The
field testing of this culvert was performed on October 16, 2006.

LAK-608-0303
The installation year of the culvert located on State Route 608 was 1961. The culvert had
the skew angle of 9 degrees. Concrete was cast at the bottom of the culvert as shown in
Figure 3.7(f). The culvert had damaged parts at the plate connections due to steel
corrosion especially at the crown around the mid length of the culvert. The field test on
August 27, 2007 was not carried out very well because of heavy rain. Most strain
measurements were useless due to noise and bad connections. However, the displacement
sensors worked well, and provided reliable data.

36
(a) HOC-093-1975

(b) HOC-664-1172

(c) JAC-327-0077

(d) KNO-013-2162 (e) KNO-062-1480

(f) LAK-608-0303

Figure 3.7: The inspected condition of the culverts before testing.

37
LIC-079-2276
The culvert shown in Figure 3.8(a) was located near Newark in Licking County, Ohio
and it was the first culvert tested on November 4, 2005 in this study. The culvert had a
skew angle of 41 degrees from the perpendicular line to the roadway, and consisted of six
corrugated plates. The inside of the culvert was clean with some rust on the bottom plates.

LIC-661-0805
The corrugated metal culvert was constructed in 1982, on State Route 661, near Granville.
This culvert with a skew angle of 9 degrees was tested on September 7, 2006. The shape
of the culvert was pipe-arch, and nine corrugated plates were bolted together. The culvert
did not have any corrosion damage except on small portions of bottom plates as shown in
Figure 3.8(b).

MAD-038-0935
MAD-038-0935 was installed in 1961, and was located on State Route 38 near London in
Madison County, Ohio. The installation direction of the culvert was 14 degrees from the
perpendicular line to the roadway. Nine corrugated plates were bolted together to make
the pipe-arch shape. Before the tests, the culvert was inspected visually. The typical
damage is shown in Figure 3.8(c). There were extensive deposits of sand and gravel
inside the culvert. Four structural plates at the bottom of the culvert were corroded
extensively. As a result of corrosion, parts of steel corrugation were lost, and several
holes were found in one corner near the bottom. Near the top of the culvert, the
connections between plates were detached and created gaps at several locations. The
metal around the gaps had turned white due to chemical reaction, such as carbonation.
This culvert was tested on July 25, 2006.

MAD-056-1490
The corrugated steel culvert built in 1956 was located on State Route 56 near London. It
was tested with MAD-038-0935 on the same day. This culvert was installed
perpendicular to the centerline of the road and consisted of seven corrugated plates to
form the pipe-arch shape as shown in Figure 3.8(d). Three structural plates at the bottom

38
of the culvert were lightly corroded. Some bolts at the crown are damaged also due to
corrosion.

MAD-665-1071
The culvert was installed in 1964, and was located on State Route 665 near Wrightsville.
It was tested on June 19, 2006. A skew angle of 15 degrees from the perpendicular line to
the roadway was the installation direction of the culvert. Eight corrugated plates were
bolted together to form the pipe-arch shape as shown in Figure 3.8(e). The bottom of the
culvert was lightly corroded. Overall observed condition of the culvert was good.

MAR-229-0093
The culvert was constructed in 1972, and tested on July 11, 2006. The location of the
culvert was on State Route 229 near Marion in Marion County, Ohio. The culvert
consisted of eight corrugated plates. According to inspection results, there were extensive
deposits of mud and sand inside the culvert. Thus, the bottom condition of the culvert
could not be checked. However, some plates on one side were corroded. Most of bolts on
the damaged plate were also corroded. A lot of bolts at the crown were lightly damaged
due to steel corrosion as shown in Figure 3.8(f).

MAR-309-1248
The corrugated steel culvert was built in 1968, and tested on October 25, 2006. The
culvert had a skew angle of 11 degrees from the perpendicular line to the roadway. Seven
corrugated plates were bolted together to make the pipe-arch shape. The appearance of
the culvert seemed to be very clean. Figure 3.8(g) shows a view of the culvert.

MAR-423-0214
The skew angle of this 35 year old culvert was 23 degrees. Seven corrugated plates were
used as can be seen in Figure 3.8(h). Bottom plates were lightly corroded. Some bolts on
each side were also damaged. This culvert was tested on July 18, 2006.

39
MAR-423-0379
The corrugated steel culvert shown in Figure 3.8(i) was constructed with MAR-423-0214
in 1972 at the same time. The experimental date of this culvert was also the same as
MAR-423-0241. Lightly damaged parts due to steel corrosion include some bottom plates
and some bolts on the side walls. Overall condition of the culvert was almost the same as
that of MAR-423-0214.

MRW-019-0456
This pipe-arch culvert was installed in 1997, and is located on State Route 19 near Mt.
Gilead in Morrow County, Ohio. The culvert is skewed with an angle of 34 degrees from
the perpendicular line to the roadway. In order to form the shape of the pipe-arch culvert,
seven corrugated steel plates were bolted together as shown in Figure 3.8(j). The inside
condition of the culvert was very clean. However, the bottom of the culvert was lightly
corroded. This culvert was tested on July 6, 2006.

MRW-095-1637
The corrugated steel culvert was located on State Route 95 near Chesterville. The culvert
was built in 1966, and tested on September 29, 2006. The skew angle between culvert
and the roadway was 5 degrees. In order to fabricate the pipe-arch shape, seven
corrugated plates were bolted together as shown in Figure 3.8(k). Bottom plates were
lightly corroded. Overall condition of the culvert was clean and good at the time of
testing.

MRW-095-1796
This 10.5 ft high, 14 ft wide and 168 ft long culvert was constructed in 1970 on State
Route 95 near Chesterville in Morrow County, Ohio. The culvert had a skew angle of 20
degrees from the perpendicular line to the roadway. Seven corrugated galvanized steel
plates were bolted along their edges to construct the pipe-arch shape. The inside of the
culvert was clean; however some of the bottom plates showed rust. Figure 3.8(l) shows
general view of the culvert during the test, which was conducted on the day MRW-095-
1637 was tested.

40
(a) LIC-079-2276 (b) LIC-661-0805

(c) MAD-038-0935

(d) MAD-056-1490 (e) MAD-665-1071 (f) MAR-229-0093

(g) MAR-309-1248 (h) MAR-423-0214 (i) MAR-423-0379

(j) MRW-019-0456 (k) MRW-095-1637 (l) MRW-095-1796

Figure 3.8: The inspected condition of the culverts before testing.

41
NOB-083-0419
This culvert was installed in 1941, and located on State Route 83 in Noble County, Ohio.
It was tested on June 21, 2007. The skew angle was 45 degrees from the perpendicular
line to the centerline of the road. Figure 3.9(a) shows the overall view as well as damaged
parts of the culvert. As it can be seen in the figure, a concrete wall was cast above each
end of the culvert up to the road level. The culvert had a lot of damage concentrations at
the plate connections due to corrosion especially near the entrance of the culvert. During
the inspection of the culvert, the concrete walls at the ends were damaged and broken
along a straight vertical crack immediately above the crown of the culvert. Water flows
into the crown through the opening of these cracks.

PIC-138-0091
The corrugated steel culvert was installed in 1981 on State Route 138 near Clarksburg in
Ross County, Ohio. The culvert had a skew angle of 2 degrees from the perpendicular
line to the roadway. Eight corrugated plates were bolted together to fabricate the pipe-
arch shape. There were extensive deposits of sand and gravel inside the culvert.
Localized bulges were found on the culvert walls. These bulges are possibly the result of
punctures to the external wall. These punctures are often caused during construction by
the rocks in the backfill soil because a lot of large size rocks were placed over the culvert.
Bolts along the crown were corroded. The condition of the culvert at the time of testing
on May 15, 2006 can be seen in Figure 3.9(b).

PIC-159-0868
The pipe culvert was set perpendicular to the roadway. The culvert consisted of six
corrugated plates. The plates were bolted together to form the circular shape. The bottom
plates of the pipe were seated into concrete as shown in Figure 3.9(c). The bottom 40
percent of the walls were also supported by concrete walls at each end of the culvert. The
inside of the pipe was very clean. However, the interface between pipe and concrete walls
and bottom was detached and had visible gaps. A lot of rust stains were found around the
gaps. This indicated that bottom plates of the culvert were damaged due to steel corrosion.

42
The bottom concrete played only a role of smooth water pathway, and was not
strengthening or supporting the culvert. This culvert was tested on September 15, 2006.

PIC-752-0748
This pipe-arch culvert was installed in 1981 and is located on State Route 752 near
Ashville. It was tested along with PIC-159-0868 on September 15, 2006. The culvert had
a skew angle of 42 degrees from the perpendicular line to the roadway. In order to form
the culvert structure, nine corrugated steel plates were bolted together. Concrete was cast
over the bottom plates of the pipe to enhance the strength of the culvert near the bottom
and to prevent damage due to corrosion. Such concrete applications appear to effectively
prevent the corrosion damage in culver bottom plates. The inside of the culvert was very
clean. The overall shape of the culvert is shown in Figure 3.9(d).

SAN-412-0575
This culvert constructed in 1933 was tested on September 10, 2007, and was the oldest
one during this study. The location of the culvert was on State Route 412 near Fremont in
Sandusky County. This culvert was installed perpendicular to the centerline of the road.
Nine corrugated plates were bolted together to fabricate the pipe-arch shape. The water
level in the culvert was high and muddy, and the bottom of the culvert could not be
inspected. The connections between the plates were detached and created gaps mainly
because of steel corrosion. Most of the bolts on the damaged plates were severely
corroded. The damaged condition of the culvert is shown in Figure 3.9(f).

SAN-510-0412
This corrugated steel culvert was constructed in 1960 on State Route 510 near Fremont.
The culvert had a skew angle of 30 degrees from the perpendicular line to the roadway.
The culvert shown in Figure 3.9(e) consisted of seven corrugated plates. Bottom
structural plates of the culvert were damaged due to steel corrosion. The connections
between these plates were detached and created gaps, which allowed for leakage of water
from backfill soil. There were extensive deposits of sand inside the culvert. However, the

43
overall condition of the culvert was not bad. This culvert was tested on September 10,
2007.

TRU-193-0215
The corrugated steel culvert was installed in 1954 on State Route 193 near Niles in
Trumbull County, Ohio. The date of the field testing was August 23, 2007. This culvert
had a skew angle of 20 degrees from the perpendicular line to the roadway. Eight
corrugated plates were bolted together to fabricate the pipe-arch shape. There were
extensive corrosion damages inside the culvert as shown in Figure 3.9(g). Four structural
plates at the bottom of the culvert were corroded. At the crown of the culvert, most of the
bolts were disintegrated or lost due to corrosion. Severe water leakage on or around the
damaged bolts at the crown was observed. Most of the plates near the entrance
experienced extensive corrosion.

44
(a) NOB-083-0419

(b) PIC-138-0091

(c) PIC-159-0868 (d) PIC-752-0748 (e) SAN-510-0412

(f) SAN-412-0575

(g) TRU-193-0215

Figure 3.9: The inspected condition of the culverts before testing.

45
CHAPTER 4

EXPERIMENTAL TEST RESULTS

4.1 Introduction
This chapter presents the field test results from the 39 culverts subjected to static
and dynamic loads. The experimental test results include culvert deflections and strains
measured by displacement sensors and strain gauges. As described in Chapter 3 (Figure
3.1), deflections at five points around the culvert, slippage at the crown between the
plates, and strains at 14 different locations on the culvert were measured during the tests.
Heavily loaded trucks were used to apply static and dynamic loads on the culvert. Table
4.1 shows the measured total weights and individual axle loads for the truck used for each
culvert test as well as maximum deflections and strains measured during the static tests.
The drop axle is the rear axle closest to the steering axle in the front, first drive axle is the
middle rear axle; and the second drive axle is the last axle in the rear of the truck (Figure
3.3). Some trucks had only two rear axles without a drop axle.
For the static tests, truck loads were applied at ten different locations above each
culvert as shown in Figure 3.3. The static loads were kept over the culvert approximately
one minute at each of the test locations to record the displacements and strains until the
measurements were stable. Dynamic load tests were conducted at six truck speeds
varying from 5 mph to 40 mph. However, the maximum speed varied depending on the
field conditions or the legal speed limit.

46
Table 4.1: Applied truck loads and maximum measured displacements and strains.

Truck weight (kips) Max. Max.


Culvert Steering Drop 1st drive 2nd drive Total deflection strain
axle axle axle axle weight (in.) (× 10-6)
ADA-041-0096 24.5 6.0 27.4 26.7 84.6 0.0109 22
ALL-309-1664 27.4 - 29.5 28.3 85.2 0.0134 23
ALL-696-0410 27.4 - 29.5 28.3 85.2 0.0236 39
ATB-193-2336 13.5 - 28.1 28.8 70.4 0.0276 34
ATB-307-1801 13.5 - 28.1 28.8 70.4 0.0150 20
CLI-068-0942 22.9 8.3 23.1 23.2 77.5 0.0008 -
COL-170-1981 14.0 - 29.9 29.1 73.0 0.0374 34
COL-172-1209 14.0 - 29.9 29.1 73.0 0.0287 30
DEF-018-1560 22.0 - 30.5 29.5 82.0 0.0142 18
DEL-037-0333 23.7 11.1 19.6 18.6 73.0 0.0028 7
FAY-734-0790 19.1 13.0 20.0 19.2 71.3 0.0487 44
FRA-062-2688 23.1 8.9 28.7 28.0 88.7 0.0201 27
GAL-218-0578 27.3 - 29.6 28.9 85.8 0.0382 43
HIG-134-0782 24.7 - 29.1 28.7 82.5 0.0449 69
HOC-093-1975 20.9 7.8 27.8 26.9 83.4 0.0382 40
HOC-664-1172 20.9 7.8 27.8 26.9 83.4 0.0276 48
JAC-327-0077 27.3 - 29.6 28.9 85.8 0.0732 77
KNO-013-2162 23.0 6.2 28.3 27.4 84.9 0.0107 22
KNO-062-1480 23.0 6.2 28.3 27.4 84.9 0.0008 -
LAK-608-0303 14.1 - 29.7 29.2 73.0 0.0551 69
LIC-079-2276 16.0 12.0 21.8 22.0 71.8 0.0173 28
LIC-661-0805 17.8 6.0 24.2 24.0 72.0 0.0091 11
MAD-038-0935 14.0 13.6 21.4 20.8 69.8 0.0319 108
MAD-056-1490 13.9 13.7 21.4 20.8 69.8 0.0085 24
MAD-665-1071 15.4 14.0 16.5 19.5 65.4 0.0273 36
MAR-229-0093 17.2 12.3 19.9 20.1 69.5 0.0160 17
MAR-309-1248 23.7 11.1 19.6 18.6 73.0 0.0126 11
MAR-423-0214 16.0 6.7 22.8 22.7 68.2 0.0072 11
MAR-423-0379 16.0 6.7 22.8 22.7 68.2 0.0191 24
MRW-019-0456 14.7 15.0 23.9 24.1 77.7 0.0162 19
MRW-095-1637 20.0 8.5 21.5 20.9 70.9 0.0156 24
MRW-095-1796 20.0 8.5 21.5 20.9 70.9 0.0083 12
NOB-083-0419 19.3 8.9 21.3 23.7 73.2 0.0350 52
PIC-138-0091 14.6 7.3 16.5 15.8 54.2 0.0135 44
PIC-159-0868 11.5 6.0 17.0 17.0 51.5 0.0035 4
PIC-752-0748 11.5 6.0 17.0 17.0 51.5 0.0043 12
SAN-412-0575 18.5 - 31.5 29.0 79.0 0.0886 89
SAN-510-0412 18.5 - 31.5 29.0 79.0 0.0472 40
TRU-193-0215 14.1 - 29.7 29.2 73.0 0.0189 41

47
Table 4.1 continued

Truck weight (kN) Max. Max.


Culvert Steering Drop 1st drive 2nd drive Total deflection strain
axle axle axle axle weight (mm) (× 10-6)
ADA-041-0096 109.0 26.7 121.9 118.8 376.3 0.277 22
ALL-309-1664 121.9 - 131.2 125.9 379.0 0.340 23
ALL-696-0410 121.9 - 131.2 125.9 379.0 0.599 39
ATB-193-2336 60.1 - 125.0 128.1 313.2 0.701 34
ATB-307-1801 60.1 - 125.0 128.1 313.2 0.381 20
CLI-068-0942 101.9 36.9 102.8 103.2 344.7 0.020 -
COL-170-1981 62.3 - 133.0 129.4 324.7 0.950 34
COL-172-1209 62.3 - 133.0 129.4 324.7 0.729 30
DEF-018-1560 97.9 - 135.7 131.2 364.8 0.361 18
DEL-037-0333 105.4 49.4 87.2 82.7 324.7 0.071 7
FAY-734-0790 85.0 57.8 89.0 85.4 317.2 1.237 44
FRA-062-2688 102.8 39.6 127.7 124.6 394.6 0.511 27
GAL-218-0578 121.4 - 131.7 128.6 381.7 0.970 43
HIG-134-0782 109.9 - 129.4 127.7 367.0 1.140 69
HOC-093-1975 93.0 34.7 123.7 119.7 371.0 0.970 40
HOC-664-1172 93.0 34.7 123.7 119.7 371.0 0.701 48
JAC-327-0077 121.4 - 131.7 128.6 381.7 1.859 77
KNO-013-2162 102.3 27.6 125.9 121.9 377.7 0.272 22
KNO-062-1480 102.3 27.6 125.9 121.9 377.7 0.020 -
LAK-608-0303 62.7 - 132.1 129.9 324.7 1.400 69
LIC-079-2276 71.2 53.4 97.0 97.9 319.4 0.439 28
LIC-661-0805 79.2 26.7 107.6 106.8 320.3 0.231 11
MAD-038-0935 62.3 60.5 95.2 92.5 310.5 0.810 108
MAD-056-1490 61.8 60.9 95.2 92.5 310.5 0.216 24
MAD-665-1071 68.5 62.3 73.4 86.7 290.9 0.693 36
MAR-229-0093 76.5 54.7 88.5 89.4 309.2 0.406 17
MAR-309-1248 105.4 49.4 87.2 82.7 324.7 0.320 11
MAR-423-0214 71.2 29.8 101.4 101.0 303.4 0.183 11
MAR-423-0379 71.2 29.8 101.4 101.0 303.4 0.485 24
MRW-019-0456 65.4 66.7 106.3 107.2 345.6 0.411 19
MRW-095-1637 89.0 37.8 95.6 93.0 315.4 0.396 24
MRW-095-1796 89.0 37.8 95.6 93.0 315.4 0.211 12
NOB-083-0419 85.9 39.6 94.7 105.4 325.6 0.889 52
PIC-138-0091 64.9 32.5 73.4 70.3 241.1 0.343 44
PIC-159-0868 51.2 26.7 75.6 75.6 229.1 0.089 4
PIC-752-0748 51.2 26.7 75.6 75.6 229.1 0.109 12
SAN-412-0575 82.3 - 140.1 129.0 351.4 2.250 89
SAN-510-0412 82.3 - 140.1 129.0 351.4 1.199 40
TRU-193-0215 62.7 - 132.1 129.9 324.7 0.480 41

48
4.2 Measurement of culvert deformations
In this study, the static and dynamic responses of 39 culverts were monitored
while the culverts were carrying the load applied by a heavily loaded truck. In general,
the overall measured response of each of the 39 culverts was similar. The main difference
was in the magnitude of the measured response quantities. Examples of experimental data
from three culverts, MRW-019-0456, MAD-038-0935 and JAC-327-0077 will be
presented and examined in this section. The measured test data for the other culverts are
provided in Appendix A.

4.2.1 Static response


Figures 4.1 through 4.3 show the complete test record, which include static and
dynamic deflections recorded by the five displacement sensors attached on each culvert
(Figure 3.1). The test truck was stopped over the culvert for one to two minutes for each
static load case (Figure 3.3). The measured static displacements greatly depend on the
location of the loading truck.
MRW-019-0456 was loaded for about 850 seconds while the truck was stopped
for about one minute for each static load case. It took another 500 seconds (between 850
and 1350 seconds in Figure 4.1) to complete the dynamic testing. Figure 4.1 shows that
the maximum deflection at the crown of the culvert MRW-019-0456 is 0.0162 in. (0.41
mm), measured during the static load case, LC 8. The vertical deflection at the crown was
measured by the displacement sensor identified as CH #03 (Figure 3.1). The deflections
near the spring line or bottom of side walls on each side (CH #01 and CH #05) are very
small for all load cases. This is an indication that side pressure resulting from a load over
pipe-arch culverts can be negligible. In case of LC 5 and LC 6, deflections of CH #02 are
larger than those of CH #03, because the loads applied by the heavy rear axles are closer
to the culvert (Figure 3.3). Similarly, deflections recorded by CH #04 start to increase as
the rear axles move over that side of the culvert (e.g., LC 9 and LC 10). Also, the
deflections measured by CH #02 for LC 5 and LC 6 are almost mirror images of those
measured by CH #04 for LC 9 and LC 10 as a result of symmetry. For most static load
cases, the maximum deflection occurs at the crown (measured by CH #03).

49
Figures 4.2 and 4.3 show the measured static and dynamic deflection histories for
culverts MAD-038-0935 and JAC-327-0077. The maximum static deflections are 0.0319
in. (0.81 mm) and 0.0732 in. (1.86 mm), respectively. Although the overall deformation
response of all culverts is similar, a few discrepancies are observed for different geometry
and environmental conditions. For example, the displacement in CH #03 or the crown
displacement for LC 10 is unexpectedly higher than that for LC 9 in Figure 4.2. This
difference of approximately 0.0016 in. (0.04 mm) may have resulted from skewed angle
of the culvert or local slip between the bolted plates.
The maximum measured deflections and deformed shapes of culvert MAD-38-
0935 for the ten static load cases are shown in Figure 4.4. The largest deflections
consistently occur at the crown and both shoulders of the culvert, which is consistent with
the location of the truck at the time of the measurements. However, the deflections near
the spring line of the culvert (measured by CH #1 and CH #5) are usually very small.
This may have resulted from a balance of increased earth pressure on the sides of the
culvert wall (deflecting in) and increased bending moment in the culvert wall due to arch
support of the vertical applied load (deflecting out).

50
0.5
LC 8 5 mph 15 mph 30 mph
CH #03 LC 9
0.4 0.015
Deflection (mm) LC 7

Deflection (in.)
40 mph
0.3 LC 10
LC 5 0.010
0.2 LC 3
LC 2 LC 6
0.005
0.1 LC 4
LC 1
0 0
10 mph 20 mph
-0.1
0 200 400 600 800 1000 1200 1400
Time (s)
0.5
CH #02
0.4 0.015
CH #04
Deflection (mm)

Deflection (in.)
0.3
0.010
0.2
0.005
0.1

0 0

-0.1
0 200 400 600 800 1000 1200 1400
Time (s)
0.5
CH #01
0.4 0.015
CH #05
Deflection (mm)

0.3 Deflection (in.)


0.010
0.2
0.005
0.1

0 0

-0.1
0 200 400 600 800 1000 1200 1400
Time (s)

Figure 4.1: Deflections in culvert MRW-019-0456 under static and dynamic loads.

51
0.9
LC 7 LC 10 10 mph 30 mph
CH #03 0.03
Deflection (mm) 0.6 LC 2 LC 8 LC 9

Deflection (in.)
LC 6
0.02

0.3 LC 5
LC 1 LC 3 0.01
5 mph

0 LC 4
0
40 mph
15 mph 20 mph
-0.3 -0.01
0 200 400 600 800 1000 1200
Time (s)
0.9
CH #02 0.03
0.6 CH #04
Deflection (mm)

Deflection (in.)
0.02

0.3 0.01

0 0

-0.3 -0.01
0 200 400 600 800 1000 1200
Time (s)
0.9
CH #01 0.03
0.6 CH #05
Deflection (mm)

0.02 Deflection (in.)

0.3 0.01

0 0

-0.3 -0.01
0 200 400 600 800 1000 1200
Time (s)

Figure 4.2: Deflections in culvert MAD-038-0935 under static and dynamic loads.

52
2.5
5 mph 15 mph
2.0 CH #03 LC 9 20 mph 0.08
10 mph
Deflection (mm) 30 mph 40 mph
1.5 0.06

Deflection (in.)
LC 8
LC 10
1.0 0.04
LC 3
0.5 LC 2 LC 4 0.02
0 LC 7 0
LC 1
-0.5 LC 5 LC 6 -0.02
-1.0
0 200 400 600 800 1000
Time (s)
2.5
2.0 CH #02 0.08
CH #04
Deflection (mm)

1.5 0.06

Deflection (in.)
1.0 0.04
0.5 0.02
0 0
-0.5 -0.02
-1.0
0 200 400 600 800 1000
Time (s)
2.5
2.0 CH #01 0.08
CH #05
Deflection (mm)

1.5 0.06
Deflection (in.)
1.0 0.04
0.5 0.02
0 0
-0.5 -0.02
-1.0
0 200 400 600 800 1000
Time (s)

Figure 4.3: Deflections in culvert JAC-327-0077 under static and dynamic loads.

53
Figure 4.4: Deformed shape and maximum deflection of culvert MAD-038-0935 for ten
static load cases (deflections are magnified 254 times).

54
4.2.2 Dynamic response
As shown in Figures 4.1 through 4.3, two displacement spikes were recorded for
each dynamic test; the first peak corresponding to the truck going over the culvert at
uniform test speed, and then the second peak for backing up for the next dynamic test.
The overall magnitudes of maximum dynamic displacements are similar to the maximum
static displacements, which suggest that the truck speed did not have a major influence on
culvert deflections. This issue will be discussed in detail later in this section.
Culvert deflections were measured during the dynamic tests for truck speeds of 5,
10, 15, 20, 30 and 40 mph. The dynamic test results for JAC-327-0077 are shown in
Figure 4.5. Each channel recorded two peak displacements while the truck was passing
over the culvert. The first peak coincides with the load applied by the steering axle of the
truck and the second peak is due to the weight on the rear axles. The larger magnitude of
the second peak is an indication of larger load on the rear axles. The maximum deflection
is essentially independent of truck speed. However, the maximum dynamic deflections
are, in general, smaller than the maximum static deflections. In Figure 4.5, the term
“dynamic loading time” indicates the time period during which deflections are changing
due to the truck passing over the culvert. As expected, the dynamic loading time (DLT)
decreases with increasing truck speed. The relationship between DLT and truck speed
will be discussed at Chapter 5 in detail.

4.3 Measurement of culvert strains


Strain histories at 14 points on each culvert were measured during the static and
dynamic tests. Channels 7 through 20 (CH #07 – CH #20 in Figure 3.2) were used to
monitor strains in the longitudinal and transverse directions. Maximum strains measured
at the culvert crowns for 39 culverts were listed in Table 4.1. In this section, examples of
selected detailed strain data for three culverts, MRW-019-0456, MAD-038-0935 and
JAC-327-0077 are introduced. Strains measured in all other culverts are provided in
Appendix B. Although most strain gauges were placed and numbered as shown in Figure
3.2, in some special cases strain gauges were attached at different locations to check the
behavior of damaged parts, other directions, or different unique locations. Section 4.3.2
describes these special cases.

55
1.6 0.06
Second peak
Speed = 5 mph (8 km/h) CH #01
1.2
Max deflection = 0.0545 in. CH #02
Deflection (mm) 0.04

Deflection (in.)
0.8 First peak CH #03
CH #04
0.4 CH #05 0.02
0
0
-0.4
-0.02
-0.8 Dynamic loading time
0 3 6 9
Elapsed time (s)

1.6 0.06
Speed = 10 mph (16 km/h) CH #01
1.2 Max deflection = 0.0517 in. CH #02
Deflection (mm)

0.04

Deflection (in.)
0.8 CH #03
CH #04
CH #05 0.02
0.4

0 0

-0.4
-0.02
-0.8
0 3 6 9
Elapsed time (s)
1.6 0.06
Speed = 15 mph (24 km/h) CH #01
1.2 Max deflection = 0.0512 in. CH #02
Deflection (mm)

0.04
Deflection (in.)
0.8 CH #03
CH #04
CH #05 0.02
0.4

0 0

-0.4
-0.02
-0.8
0 3 6 9
Elapsed time (s)

Figure 4.5: Deflections of JAC-327-0077 for truck speeds varying from 5 to 40 mph (8 to
64 km/hr).

56
Figure 4.5 continued

1.6 0.06
Speed = 20 mph (32 km/hr) CH #01
1.2 Max deflection = 0.0517 in. CH #02
Deflection (mm)

0.04

Deflection (in.)
0.8 CH #03
CH #04
CH #05 0.02
0.4

0 0

-0.4
-0.02
-0.8
0 3 6 9
Elapsed time (s)
1.6 0.06
Speed = 30 mph (48 km/h) CH #01
1.2 Max deflection = 0.0495 in. CH #02
Deflection (mm)

0.04

Deflection (in.)
0.8 CH #03
CH #04
CH #05 0.02
0.4

0 0

-0.4
-0.02
-0.8
0 3 6 9
Elapsed time (s)
1.6 0.06
Speed = 40 mph (64 km/h) CH #01
1.2 Max deflection = 0.0516 in. CH #02 0.04
Deflection (mm)

Deflection (in.)

0.8 CH #03
CH #04
CH #05 0.02
0.4

0 0

-0.4
-0.02
-0.8
0 3 6 9
Elapsed time (s)

57
4.3.1 Static response
The maximum strain measured in MAD-038-0935 is 108 μs, and is the largest
strain recorded in this study. The low appraisal number for this culvert (N = 3) reflects its
deteriorating condition, which likely produced the larger strains, hence stresses during
testing. Figure 4.6 presents the strains measured at 14 locations on culvert MAD-038-
0935. The strain gauge locations on this culvert were the same as those shown in Figure
3.2.
In order to check whether strain gauges worked well, two strain histories. CH #07
and CH #10 which are located in the peak of the corrugation at the crown are compared.
As it can be seen in Figure 4.6, the strain variations in channels CH #07 and CH #10 are
very similar. Although there is a small difference in strain magnitudes, the overall
response is very close and captured well. This means that the strain gauges attached at the
crown of the culvert worked well and the associated data were reliable. Thus, only CH
#07 is used for the analysis of culvert behavior at the crown. Similar verifications were
sought in other culverts by comparing the data from similar strain gauges and by
comparing the strain histories measured in the two loading directions.
Three strain histories, measured at channels CH #07, CH #08 and CH #09, are
shown in Figure 4.6. The objective is to compare the corrugation behavior under the same
loading at different locations; one from CH #07 at the peak of the corrugation, another
from CH #09 at the valley of the corrugation, and the other from CH #08 in the middle
of corrugation between CH #07 and CH #09 at the crown. As shown in the figure, CH
#09 (in the valley) was mainly in compression while the strains measured at CH #07 (at
the peak) were predominantly in tension due to flexure of the corrugated plate (e.g.,
similar to a beam under transverse loading). However, the absolute magnitudes of tensile
and compressive strains are not the same since the neutral axis is not located at mid-
height of the corrugation. CH #08 showed approximately average values of CH #07 and
CH #09. However, CH #08 is not in the middle of corrugation between CH #07 and CH
#09. This is partially because the culvert plates are curved and the soil pressure is
probably not distributed uniformly. Some researchers have used beam theory to convert
strains to thrust forces and moments (Beal 1982, Chelliah 1992). Our investigation,
however, has indicated that more accurate analytical methods and better simplification

58
are needed to interpret strain measurements and to calculate thrusts and moments in the
culvert plates. For example, the measured strains near the spring line of the culvert are
very small for all load cases (i.e., CH #18), while the thrust forces are large. In general,
the measured strains, as an indication of stresses in the culvert, seem to be proportional
with the culvert deflections. In addition, the strains generally increase as the deflections
increase. Figures 4.7 and 4.8 show the measured static and dynamic strain histories for
culverts MRW-095-1796 and JAC-327-0077. Overall strain response of these and other
culverts is similar.

4.3.2 Variation in strain gauge locations in different culverts


In this section explained are some exceptional culverts in which strain gauges
were attached at different locations. The strain gauge locations, which differ from the
locations shown in Figure 3.2, are numbered from CH #21. The culvert LIC-079-2276
was the first culvert tested in this research. In order to find critical parameters and
optimal locations to investigate culvert behavior, eight strain gauges were attached in the
longitudinal direction as shown in Figure 4.9. The measured strain histories are shown for
the culvert in Figure B.20 in Appendix B. From the test results, it was found that strains
in the circumferential direction are more important than longitudinal ones. Similarly, the
strains measured at the crown (CH #07 and CH#09) differed due to corrugation. The
second test culvert had damage in some parts of the culvert. There were some bulges and
deflection at the crown of this culvert, PIC-138-0091. Five strain gauges were attached
on the bulges to investigate the behavior of damaged parts. Figure B.30 shows the
recorded strain histories. Strain histories recorded in channels CH #21 to CH #25 exhibit
the behavior of damaged parts and give larger values compared with other critical
locations where CH #07, CH #08, and CH #09 were attached.

59
80

Strain (μs) 40

-40

-80
CH #07
-120 CH #10

80
CH #07
40 CH #08
CH #09
Strain (μs)

-40

-80

-120

80
CH #13
CH #16
40
Strain (μs)

-40

-80

-120

0 200 400 600 800 1000 1200


Time (s)

Figure 4.6: Measured strain versus time plot for culvert MAD-038-0935.

60
Figure 4.6 continued

80
CH #16
40 CH #18
Strain (μs)

-40

-80

-120

80
CH #14
40 CH #17
Strain (μs)

-40

-80
Longitudinal direction
-120

100
CH #11
50 CH #19
CH #20
Strain (μs)

-50

-100

-150
0 200 400 600 800 1000 1200
Time (s)

61
10
CH #10
CH #12
Strain (μs) 5

-5

-10

-15
10
CH #07
CH #08
5 CH #09
Strain (μs)

-5

-10

-15
10
CH #13
5 Data for CH #15 is not available due to noise
Strain (μs)

-5

-10

-15
0 200 400 600 800 1000 1200
Time (s)

Figure 4.7: Measured strain versus time plot for culvert MRW-095-1796.

62
Figure 4.7 continued

10
CH #16
CH #18
5
Strain (μs)

-5

-10

-15
10
CH #14
5 CH #17
Strain (μs)

-5

-10 Longitudinal direction

-15
10
CH #11
5 CH #19
CH #20
Strain (μs)

-5

-10

-15
0 200 400 600 800 1000 1200
Time (s)

63
100
CH #10
CH #12
Strain (μs) 50

-50

-100
100
CH #08 Data for CH #07 is not available due to noise
CH #09
50
Strain (μs)

-50

-100
100
CH #13

50 Data for CH #15 is not available due to noise


Strain (μs)

-50

-100
0 200 400 600 800 1000
Time (s)

Figure 4.8: Measured strain versus time plot for culvert JAC-327-0077.

64
Figure 4.8 continued

100
CH #16
CH #18 No data available
50
Strain (μs)

-50

-100
100
CH #14

50 Longitudinal direction
Strain (μs)

Data for CH #17 is not available due to noise


0

-50

-100
100
CH #19
CH #20
50
Strain (μs)

-50

-100
0 200 400 600 800 1000
Time (s)

65
Figure 4.9: Strain gauge locations of the culvert LIC-079-2276.

66
CHAPTER 5

EVALUATION TEST RESULTS

5.1 Introduction
Many researchers have reported experimental culvert performance during and
after construction (see Section 2.1). Although extensive experimental research has been
conducted by testing corrugated metal culverts in the field, there is not sufficient
experimental evidence to evaluate the existing culvert response considering the effect of
related parameters such as culvert span, culvert height, backfill height, nature and
magnitude of loading on the surface, and maximum deflection and maximum strain
experienced by the culvert. This chapter compares and evaluates the experimental
response of culverts presented in Chapter 4. The aim of this chapter is to investigate the
relationship between two or more parameters such as culvert span, culvert height, backfill
height, truck loading, maximum deflection, and maximum strain obtained from 39
experimental tests. In this chapter, live load models are introduced to represent the
individual truck wheel loads as uniformly distributed or equivalent live load based on
elastic solutions, i.e., Boussinesq theory and AASHTO method using a spreading factor.

5.2 Relationship between the factors affecting culvert behavior


Figures 5.1 and 5.2 show the relation between maximum deflection versus culvert
span and height, respectively. As it can be seen in these figures, the data points are
largely scattered for the 39 culverts tested in this research. There are no clear trends. In
general, the bigger the culvert size, the larger the deflections are. The data indicate that
additional factors, such as backfill height and loading, should be considered to obtain
more meaningful and accurate relationship to represent culvert behavior.
Plots of maximum deflection and maximum compressive normal strain versus
backfill height are shown in Figure 5.3. The deflection and strain values decrease

67
nonlinearly with increasing backfill height. Relatively large deflections and strains were
measured for backfill heights less than 6.5 ft (2.0 m), which indicates that backfill height
is a critical parameter affecting the response. According to AASHTO Specifications
(2000), the minimum cover for plate pipe or arch structures is 1/8 of the span length but
not less than 1.0 ft (0.3 m). All culverts tested in this study satisfy this requirement.
Maximum measured strains seem to be generally proportional to the culvert deflections
and show the same trend with regard to backfill height. For deep culverts with backfill
heights larger than 13.0 ft (4.0 m), maximum deflections were very close to zero. Thus,
the imposed truck loads had an insignificant effect on the deep culverts in this study.
Figure 5.4 shows a plot of maximum measured deflection versus total truck
weight used during tests. In general, deflections increase with increasing truck weight as
it is expected. However, there is considerable scatter in this plot, which suggests that
additional critical factors such as backfill height and culvert geometry must be considered
to establish a clearer trend for deformations in existing culverts.
Figure 5.5 presents a plot of maximum deflection versus general appraisal number
N, where low numbers indicate poor condition and high number indicate excellent
condition of a culvert. Although this plot displays considerable scatter, maximum
deflections generally decrease with improving culvert condition, i.e., with increasing N.
Indications of poor conditions would include missing bolts at joints of corrugated steel
plate joints and rusted out sections of plates, both of which would be expected to
contribute to larger deflections. Again, general appraisal number alone is not a good
indication of how a culvert behaves and deflects. It is not even clear whether or how
important the appraisal number is in determining the maximum culvert deflection.
Further analysis will be carried out to identify the significance of culvert parameters on
culvert deformation.
Figures 5.1 through 5.5 suggest that none of the parameters has a direct
relationship with the culvert deformations when they are considered one by one. Clearly,
certain parameters (e.g., backfill height) affect the response significantly, while others
(e.g., general appraisal number) may have less influence on the overall response. In the
next section, the effect of combination of two most significant parameters will be

68
discussed. The equivalent line load includes the effect of backfill height and truck live
load.

Span length (ft)


10 12 14 16 18 20 22 24
2.5
Maximum deflection (mm)

2.0 0.08

Maximum deflection (in.)


1.5 0.06

1.0 0.04

0.5 0.02

0 0
3 4 5 6 7
Span length (m)

Figure 5.1: Maximum culvert deflection versus culvert span length.

Culvert height (ft)


7 8 9 10 11 12 13
2.5
Maximum deflection (mm)

2.0 0.08 Maximum deflection (in.)

1.5 0.06

1.0 0.04

0.5 0.02

0 0
2.0 2.5 3.0 3.5 4.0
Culvert height (m)

Figure 5.2: Maximum culvert deflection versus culvert height.

69
Backfill height (ft)
0 5 10 15 20 25
2.5

Maximum deflection (mm)


0.08

Maximum deflection (in.)


2.0

1.5 0.06

1.0 0.04

0.5 0.02

0 0
0 1 2 3 4 5 6 7 8
Backfill height (m)

(a)

Backfill height (ft)


0 5 10 15 20 25
120

100
Maximum strain (10 )
-6

80

60

40

20

0
0 1 2 3 4 5 6 7 8
Backfill height (m)

(b)

Figure 5.3: Measured response for 40 corrugated metal culverts: (a) maximum deflection
and (b) maximum compressive strain versus backfill height.

70
Total weight of testing truck (kips)
50 55 60 65 70 75 80 85 90
2.5

Maximum deflection (mm)


2.0 0.08

Maximum deflection (in.)


1.5 0.06

1.0 0.04

0.5 0.02

0 0
240 280 320 360 400
Total weight of testing truck (kN)

Figure 5.4: Maximum culvert deflection versus total truck weight

2.5
Maximum deflection (mm)

2.0 0.08

Maximum deflection (in.)


1.5 0.06

1.0 0.04

0.5 0.02

0 0
0 2 4 6 8 10
General appraisal number

Figure 5.5: Maximum culvert deflection versus general appraisal number

71
5.3 Equivalent line load
In order to consider the effect of point loads (wheel loads) on the behavior of
culvert, some researchers (Bakht, 1981; Adbel-Sayed and Bakht, 1982) have used the
American Association of State Highway and Transportation Officials (AASHTO) and
Boussinesq methods to calculate a two-dimensional “equivalent line load” q that has the
same maximum stress as the imposed three-dimensional load and thus can allow the
problem to be analyzed using a two-dimensional approach. The simplest method for
estimating an equivalent line load is based on an assumed spreading factor. To compute
vertical stresses at a given depth, the surface load is assumed to spread to be effective
over a length equal to the depth times the spreading factor. Bakht (1981) have discussed
the appropriate value for this spreading factor, and reported that different spreading
factors should be used for the parallel or perpendicular directions of the culvert. This is
generally not considered when computing equivalent line loads for a plane strain or two-
dimensional finite element analysis. Currently, AASHTO assumes that a point load at the
ground surface for any depth greater than 2 ft (0.6 m) acts on a square area with length of
sides 1.75 times the depth of H. The AASHTO spreading factor of 1.75 is generally used
by other researchers to convert a point load to an equivalent line load (Powell, 1987).
Live loads can be modeled using two-dimensional plane strain finite element
analysis programs like CANDE (Culvert ANalysis and DEsign) code. According to the
assumption of plane-strain method, loads in the longitudinal direction of culvert are the
same and constant because a unit length of the culvert in the longitudinal direction is
modeled. This assumption gives reasonable results for dead loads like soil weight and
overburden loads. However, live loads over the culvert can spread in both transverse and
longitudinal direction. Another problem is that this technology can not consider a layered
media. Beal (1986) employed this technique but also pointed out that there is not an
available elasticity solution for a layered media.
Katona et al. (1976) represent concentrated wheel loads as equivalent line loads
using the Boussinesq theory as shown in Figure 5.6, and Duncan (1979) devises a
simplified equation for computing a line load for design studies using finite element
analyses. This simplified formula is a function of the axle loads and the depth of cover
over the culvert. The equation is based on elasticity solutions for an HS-20 vehicle. The

72
equivalent line load is intended to produce the same maximum vertical soil stresses at the
height of the crown as the wheel loads on the ground surface. Thus, these studies on
equivalent line loads have shown that live loads are not distributed equally in both
directions. Bakht (1982) also pointed out that a concentrated load disperses more rapidly
in the transverse direction of the culvert than in the longitudinal direction.

Figure 5.6: Idealization of a point load to a line load (Katona et al., 1976).

In this study, the AASHTO method was applied to convert each axle load Q to an
equivalent live load q. Wheel loads are assumed to be uniformly distributed at the level of
the culvert crown and thus, for multiple wheel loads, the following equation can be used.

Q
q= (5.1)
1.75 H + DW + 2W

where W is the width of each pair of wheels, and Dw is the clear distance between the
wheels (Figure 5.7).
Figure 5.8 presents the maximum measured deflections versus equivalent line
loads calculated using Equation 5.1. Compared to Figures 5.3a and 5.4, a better trend is
observed, in which the maximum deflection increases with increasing q. However,
additional scatter is observed in the data for q ≥ 2.0 kips/ft (30 kN/m). This suggests that

73
it is more difficult to predict culvert behavior when backfill height over the culvert crown
is shallow and external loads are larger. This scatter is likely due to other variables, such
as culvert age, culvert properties, general appraisal number, and differences in backfill
soil conditions.

5.4 The effect of truck speed


In Section 4.2.2, the term “dynamic loading time” (DLT) was defined as the time
required for the truck to pass over the culvert. DLT is calculated using the recorded
displacement data from CH #03. DLT changes with truck speed, culvert span length, and
possibly the skew angle of the road. Figure 5.9 shows measured ratio of DLT to
maximum DLT versus truck speed. For each culvert, DLT ratios decrease as the truck
speed increases with the maximum DLT value occurring at 5 mph (8 km/h). DLT values
at 20 mph (32 km/h) decrease by 50% to 80% as compared to values at 5mph (8 km/h).
Interestingly, however, the decreasing trend of this plot is not linear. Higher truck speeds
have progressively less effect on the DLT value, which results from the overall dynamic
response of the soil/structure system.
Figure 5.10 shows ratios of maximum dynamic culvert deflection to maximum
static deflection as a function of truck speed for all dynamic tests. These maximum
deflections primarily occurred at the crowns of the culverts (recorded by CH #03). A few
culverts yielded maximum deflections at CH #02 or CH #04, possibly due to extreme
highway slope or skew angle. With the exception of PIC-752-0748 at low speeds,
maximum dynamic deflections are consistently less than the corresponding static
deflections. The deflection ratio varied within the range from 0.7 to 0.9 for most culverts
and speeds. In addition, most culverts yielded deflection ratios that slightly decrease with
increasing truck speed, whereas other culverts show essentially constant deflection ratio
for all dynamic tests. This could be due to several factors including the structural
condition and dynamic properties (e.g., natural frequency) of the culvert. The reason for
the discrepancy of deflection ratios for PIC-752-0748 is unclear. Figure 5.10 suggests
that developing and calibrating culvert design methodologies using static deflection data
is conservative. This finding is important because it eliminates the need to perform
dynamic load tests in the field.

74
Figure 5.7: Equivalent load distribution due to wheel loads over a culvert.

Equivalent line load (kips/ft)


0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
2.5
Maximum deflection (mm)

2.0 0.08
Maximum deflection (in.)
1.5 0.06

1.0 0.04

0.5 0.02

0 0
0 10 20 30 40 50 60
Equivalent line load (kN/m)

Figure 5.8: Maximum culvert deflection versus AASHTO equivalent line load.

75
Truck speed (mph)
5 10 15 20 25 30 35 40
1.0

0.8

DLT / Max DL
0.6

0.4

0.2

0.0
0 10 20 30 40 50 60 70
Truck speed (km/h)

Figure 5.9: Normalized dynamic loading time ratio versus truck speed.

Truck speed (mph)


5 10 15 20 25 30 35 40
1.2
PIC-752-0748
1.0
Max Δdynamic / Max Δstatic

0.8

0.6

0.4

0.2

0.0
0 10 20 30 40 50 60 70
Truck speed (km/h)

Figure 5.10: Maximum deflection ratio versus truck speed (Δ denotes deflection).

76
5.5 Internal forces in plates
This section presents the bending moments and thrust forces calculated from plate
strains measured during experimental tests. Corrugated steel plate is typically treated as a
thin plate and the arch structure is crudely modeled as a simple beam with the relevant
beam forces of axial thrust and bending moments. In order to convert strains to desired
stresses, some researchers (Chelliah, 1992; Rauch, 1990; Beal, 1982) used stress-strain
relationship derived from plane-stress theory as shown in Equation 5.2:
E
σc = (ε c −νε l ) (5.2)
1 −ν 2
where σ c is the circumferential or transverse stress, E is the modulus of elasticity of the

corrugated steel, ν is Poisson’s ratio, ε c is circumferential strain, and ε l is longitudinal


strain. However, this equation is not used properly. As shown in Figure 5.11, the vertical
strain ( ε v ) instead of ε l in Equation 5.2 shall be used. In addition, they compared the
results using the plane-stress theory with predicted moments and thrusts using the plane-
strain theory. Timoshenko and Goodier (1970) recommended the plane-strain theory for
calculation of stresses in culverts and tunnels. However, the plane-strain theory cannot be
used directly. Because, in practice, vertical strains of the cross section of corrugated plate
cannot be measured, and longitudinal strains are useless due to the plane-strain
assumption that longitudinal strain is equal to zero. Thus, in this study, a generalized
plane-strain method was employed to covert strains to stresses. Equations for the
generalized plane-strain method are shown below.
εl = εo ≠ 0 (5.3)

E
σc = [(1 +ν )ε c +ν (ε v + ε o )] (5.4)
(1 +ν )(1 − 2ν )
where ε o is a constant strain along longitudinal direction, and ε v is the vertical strain

inside the corrugation. It is assumed that ε v is equal to zero because the thickness of
corrugation is very small.

77
Figure 5.11: Plane stress state in a cross section of corrugated plate.

As discussed earlier, the corrugated plate is treated as a simple thin plate in a


beam subjected to bending moment and thrust force. The resulting stresses can be
calculated from Equation 5.5.
T M ⋅c
σc = ± (5.5)
A I
where M is bending moment, T is axial thrust, A is area of cross section, I is moment
of inertia, and c is distance from the location where strain is measured to the neutral axis
as shown in Figure 5.12. From Equations 5.4 and 5.5, the following relationships can be
derived to compute the moment and thrust in the plate.
(σ 1 − σ 2 )I
M= (5.6)
c1 + c2
(σ 1c2 + σ 2 c1 ) A
T= (5.7)
c1 + c2

where c1 is distance from neutral axis to the furthest point in tension zone, c2 is distance

from neutral axis to the furthest point in compression zone, and σ 1 and σ 2 are
circumferential stresses in tension and compression zone, respectively. For the 39
culverts tested in this research, the maximum moment and thrust forces for load case 8
(See Figure 3.3) are calculated form Equations 5.6 and 5.7. The stresses in Equations 5.6
and 5.7 are calculated from Equation 5.4, where E = 30 × 106 lb/in.2 ( 2.1×108 kN/m 2 ),

78
ν = 0.3 , and ε c and ε o are listed in Table 5.1. Moment of inertia I and area A were
reported in Table 3.2. The maximum measured strains, hence, the maximum moment and
thrusts occur at the crown of the test culverts. The calculated maximum moment and
thrusts are shown in Table 5.2. The measured strains were extremely small in the deep
culverts with no reported internal forces (i.e., CLI-068-0942, and KNO-062-1480). Some
culvert strains at the crown for LC 8 are not measured because of noise, testing set-up
problem, or environmental conditions like weather (marked as “-” in Table 5.1).
Calculated maximum thrust and moment at the crown are listed in Tables 5.1. A
consistent sign convention is employed throughout this study. Positive bending moments
mean that the plate is bent downwardly, and positive axial thrusts are tensile. Although
the maximum strain in Table 5.2 occurs at the crown of the culvert MAD-038-0935, the
calculated maximum stress is 3940 lb/in.2 (27165 kN/m2) for the culvert SAN-412-0575
because of the longitudinal strain ( ε o ). Thus, ε o is one of the important factors to
determine stresses based on the plain-strain theory.

Figure 5.12: Geometry to perform data analysis for culvert forces.

79
Table 5.1: Calculated stresses based on the plane-strain theory.

ε in ε in σ1 σ2
Culvert name c c εo 2 2 2
tension compression lb/in. kN/m lb/in. kN/m2
ADA-041-0096 - - - - - - -
ALL-309-1664 8 -23 1 340 2347 -912 -6285
ALL-696-0410 38 -39 -4 1465 10103 -1644 -11337
ATB-193-236 - - - - - - -
ATB-307-1801 4 -20 0 162 1114 -808 -5569
CLI-068-0942 0 0 0 0 0 0 0
COL-170-1981 20 -34 -3 756 5211 -1425 -9825
COL-172-1209 7 -30 -9 127 875 -1367 -9427
DEF-018-1560 15 -18 -4 537 3699 -796 -5489
DEL-037-0333 2 -7 -9 -75 -517 -438 -3023
FAY-734-0790 41 -37 -14 1413 9745 -1737 -11973
FRA-062-2688 -3 -29 6 -17 -119 -1067 -7359
GAL-218-0578 31 -31 -3 1200 8274 -1304 -8990
HIG-134-0782 44 -69 -2 1742 12013 -2821 -19451
HOC-093-1975 22 -40 1 906 6245 -1598 -11018
HOC-664-1172 19 -40 -12 560 3858 -1823 -12570
JAC-327-0077 70 -90 0 2827 19491 -3635 -25060
KNO-013-2162 -8 -13 0 -323 -2228 -525 -3620
KNO-062-1480 0 0 0 0 0 0 0
LAK-608-0303 - - - - - - -
LIC-079-2276 - - - - - - -
LIC-661-0805 -3 -28 -5 -208 -1432 -1217 -8393
MAD-038-0935 108 -50 -80 2977 20525 -3404 -23469
MAD-056-1490 - - - - - - -
MAD-665-1071 19 -36 10 940 6484 -1281 -8831
MAR-229-0093 2 -17 -4 12 80 -756 -5211
MAR-309-1248 0 -5 -7 -121 -835 -323 -2228
MAR-423-0214 2 -10 -2 46 318 -438 -3023
MAR-423-0379 12 -24 -7 363 2506 -1090 -7518
MRW-019-0456 2 -19 -5 -6 -40 -854 -5887
MRW-095-1637 4 -24 0 162 1114 -969 -6683
MRW-095-1796 -2 -12 4 -12 -80 -415 -2864
NOB-083-0419 21 -46 -20 502 3461 -2204 -15195
PIC-138-0091 2 -15 -18 -231 -1591 -917 -6325
PIC-159-0868 - - - - - - -
PIC-752-0748 -2 -5 -1 -98 -676 -219 -1512
SAN-412-0575 89 -43 20 3940 27168 -1390 -9586
SAN-510-0412 38 -18 10 1708 11774 -554 -3819
TRU-193-0215 19 -41 0 767 5290 -1656 -11416

80
Table 5.2: Maximum bending moment and axial thrust in the 39 test culverts.

Bending moment Axial thrust


Culvert name
(lb-in./in.) (kN-m/m) (lb/in.) (kN/m)
ADA-041-0096 - - - -
ALL-309-1664 72 0.319 -109 -19
ALL-696-0410 153 0.683 -81 -14
ATB-193-236 0 0.000 0 0
ATB-307-1801 55 0.247 -112 -20
CLI-068-0942 0 0 0 0
COL-170-1981 125 0.555 -146 -26
COL-172-1209 86 0.380 -204 -36
DEF-018-1560 38 0.170 -26 -5
DEL-037-0333 18 0.080 -66 -12
FAY-734-0790 155 0.692 -99 -17
FRA-062-2688 47 0.207 -127 -22
GAL-218-0578 163 0.724 -100 -18
HIG-134-0782 225 1.002 -213 -37
HOC-093-1975 124 0.550 -128 -22
HOC-664-1172 118 0.523 -191 -33
JAC-327-0077 370 1.645 -277 -49
KNO-013-2162 10 0.044 -101 -18
KNO-062-1480 0 0 0 0
LAK-608-0303 - - - -
LIC-079-2276 - - - -
LIC-661-0805 45 0.199 -162 -28
MAD-038-0935 283 1.258 -146 -25
MAD-056-1490 - - - -
MAD-665-1071 98 0.438 -70 -12
MAR-229-0093 44 0.195 -119 -21
MAR-309-1248 9 0.040 -49 -9
MAR-423-0214 21 0.096 -48 -8
MAR-423-0379 64 0.287 -97 -17
MRW-019-0456 31 0.138 -81 -14
MRW-095-1637 50 0.223 -100 -18
MRW-095-1796 18 0.080 -50 -9
NOB-083-0419 134 0.594 -247 -43
PIC-138-0091 20 0.087 -79 -14
PIC-159-0868 - - - -
PIC-752-0748 5 0.024 -34 -6
SAN-412-0575 236 1.051 175 31
SAN-510-0412 129 0.576 95 17
TRU-193-0215 139 0.617 -182 -32

81
CHAPTER 6

NUMERICAL MODELING AND ANALYSIS OF CULVERTS

6.1 Introduction
Numerous researchers have used finite element computer programs to help
interpret the observed behavior of culverts or to verify the measured experimental results
(Katona, 1978; Katona et al., 1979; Chang et al., 1980; McVay, 1982; Musser, 1989;
Rauch, 1990; Webb, 1999; Kunecki and Ebeltoft, 2007). Most of the researchers have
used CANDE (Culvert ANalysis and DEsign) computer program (Katona, 1976) to
model culvert structure and to simulate soil-structure interaction, interface slip between
culvert wall and surrounding soil, and stress-strain relationship for soil. However, as
discussed in Chapter 5, two-dimensional plane-strain method, which is adopted by
CANDE, has a problem in modeling of live loads. Some researchers have analyzed
culvert behavior with three-dimensional finite element programs (Moore and Taleb,
1994; Moore and Taleb, 1999; Kerh and Yee, 2000). They showed that predictions for
culvert response under live load were generally slightly overestimated compared to those
measured in the field. However, Moore and Taleb (1999) reported that the predicted
response obtained from three-dimensional analysis program is much closer than that
obtained from conventional plane-strain analysis, which uses two-dimensional equivalent
loads to represent live loads.
This chapter presents the results from analysis of representative test culverts
modeled using CANDE-89 program. The culvert behavior predicted by CANDE is then
compared to that measured in the experimental test.

6.2 CANDE program and modeling


CANDE is a two-dimensional finite element computer program based on plane-
strain theory, which assumes that displacement in the longitudinal direction of culvert is

82
prevented, structure and soil properties are not changed during analysis, and loading is
represented as a line load along longitudinal direction. These assumptions lead to some
difficulties, for example, modeling of point loads, and bending in the longitudinal
direction. In order to solve these problems, the equivalent line method based on
AASHTO (2002) was introduced (see section 5.3). The CANDE program is used to
simulate the behavior of corrugated culverts. The code includes the effect of various
factors such as loading conditions, soil type, soil properties, culvert properties and
geometry. It can use several elastic soil models such as linear elastic, overburden
dependent, and hyperbolic.
CANDE has three solution levels: Level 1 is the elastic plate solution suggested
by Burns and Richard (1964, 1965), and Level 2 and 3 are a finite element solution with
an automated mesh generation routine and with a user-defined mesh, respectively. Level
1 cannot be used for large-span culverts because it is applicable to circular pipes deeply
buried in homogeneous soil. Usage of Level 2 is relatively easy and fast because the user
can select one of the standard meshes already defined and provided in the CANDE
database. However, the user needs to be reminded that Level 2 uses only half mesh based
on the assumption of symmetry about vertical axis of the culvert. Level 3 requires the
user to define any user-defined materials and properties, and make any geometrical shape.
In this study, most culvert analyses were performed by using user-defined meshes (Level
3). Level 3 is more applicable to create any geometrical shape as well as soil material
zones and culvert structural properties. In addition, un-symmetric load application can be
used.
Three models with 84, 183 and 339 nodes are used to evaluate the effect of mesh
resolution on the analysis results (Figure 6.1). Half models are used due to maximum
allowable bandwidth of 120 and allowable word length of 8000 to define culvert
geometry, boundary conditions, and material properties in the CANDE program. Figure
6.2 shows that the crown displacements calculated from Mesh 2 and Mesh 3 are very
close. The displacements calculated using the coarse mesh (Mesh 1) were not accurate. In
this study, a mesh resolution similar to that of Mesh 2 is selected for the analysis.

83
(a) Mesh 1 (b) Mesh 2 (c) Mesh 3

Figure 6.1: Half culvert model with 84, 183 and 339 nodes

Distance from the crown (m)


0 1 2 3 4 5
1.0

0.8
Normalized displacement

0.6

0.4
Mesh 3
Mesh 2
0.2
Mesh 1

0
0 50 100 150 200
Distance from the crown (in.)

Figure 6.2: Effect of mesh size on analysis results.

84
6.3 Finite element models of corrugated metal culverts
6.3.1 Culvert modeling
Bilinear steel material is used to represent corrugated metal model as shown in
Figure 6.3. Default material parameters for beam elements recommended for standard
corrugated steel culverts are used in this research. Elastic modulus of pipe materials ( E1 )

is 30 ×106 lb/in.2 ( 2.1× 108 kN/m 2 ), and elastic modulus of upper portion of bilinear
stress-strain curve ( E2 ) is 0. Yield stress and Poisson’s ratio of pipe material are

33, 000 lb/in.2 ( 2.3 ×106 kN/m 2 ), and 0.3, respectively.


Culvert model shown in Figure 6.4 is constructed of beam elements, which have
two external nodes with three degrees of freedom at each node, including vertical and
horizontal displacements, and rotation. The culvert plate is simulated by using a series of
straight beam-column elements, which incorporate both bending and thrust loads in the
culvert wall. The number of beam-column elements varied from 42 to 48 to model the
test culverts in this research.
Some researchers reported that the initial shape of the culvert could affect the
finite element analysis results (Katona, 1978; Boulanger et al., 1989; McGrath, 1998).
The true shapes of the corrugated test culverts were determined from field measurements.
There were not significant differences between the measured shape and specified shape
of the culverts tested in this research except for a few culverts. Thus, the specified shapes,
instead of actual measured shapes, were used for the analysis.

Figure 6.3: Bilinear stress-strain relationship used for corrugated metal.

85
Figure 6.4: Beam elements used to model a corrugated metal culvert.

6.3.2 Soil and pavement modeling


Various built-in soil models are available to represent soil elements in CANDE
program. Soil can be modeled using continuum triangular and quadrilateral elements,
which have three or four external nodes with two degrees of freedom at each node. In this
study, only two-dimensional, isoparametric, and quadrilateral elements are used to model
soil elements as shown in Figure 6.5. The mesh utilizes roller-type boundary conditions
along the vertical sides of the mesh and hinge-type boundary conditions along the bottom
of the mesh.
In this study, the hyperbolic soil model proposed by Wong and Duncan (1974)
and Duncan et al. (1980) was chosen for representing the stress-strain behavior of soil
elements because many researchers (McVay and Selig, 1982; Leonards et al., 1985;
Rauch, 1990; Webb, 1999) concluded that the hyperbolic model is the best constitutive
model available in CANDE for soils. The hyperbolic model is introduced in Appendix C.
In order to use the model, eight parameters, which can be obtained from triaxial soil test,
need to be determined. Although soil properties are critical to simulate the behavior of
the backfill material using Duncan’s hyperbolic constitutive model, soil tests are not
performed in this research due to its limited scope. Backfill material properties
recommended by ODOT’s Construction and Material Specifications (from 1951 to 1989),

86
research documents (Sargand et al., 2008; Chelliah, 1992; Rauch, 1990) are considered in
this research. Along with these recommendations, two types of backfill materials, silty
sand (SM) and, silty clayey sand (SC), are selected. The modeling parameters for these
soils are recommended by Duncan (Musser, 1989). The parameters are listed in Table 6.1
and are discussed in Appendix C.
The pavement layer was represented by quadrilateral elements, which have four
external nodes with two degrees of freedom at each node. A linear elastic stress-strain
relationship is used to represent the pavement behavior. Material properties for the
pavement suggested by Chelliah (1992) are used. Elastic modulus of pavement material
is 1.0 ×106 psi ( 6.9 ×106 kN/m 2 ), and Poisson’s ratio is 0.41.

Pavement

Culvert

Soil

Figure 6.5: Finite element model for corrugated metal culverts.

87
Table 6.1: Soil parameters used for the hyperbolic soil model in culvert analysis.

Parameters Silty clayey sand (SC) Silty sand (SM)


K 400 600
n 0.6 0.25
Rf 0.7 0.7
c 0.5 kip/ft2 (23.9 kN/m2) 0.0
φo (degree) 33 36
Δφ (degree) 0 8
Kb 200 450
m 0.5 0.0

6.3.3 Simulation of live load effect


The most difficult part of two-dimensional plane-strain finite element analysis is
the modeling of point loads on the surface. The best way to solve this problem is to
represent a point load as an equivalent line load on the culvert as explained in Section 5.3.
The equivalent live load can be defined as the line load, which produces the same vertical
pressure at the crown (for the two-dimensional plane strain analysis) as the point load on
the ground surface (for three-dimensional analysis). The calculation of equivalent line
load is necessary for CANDE program because the program is based on plane-strain
assumption. The equivalent line load for each testing truck is calculated from Equation
5.1. A tire contact area of 10 by 20 in. (254 by 508 mm) is used in the equation. The truck
geometry, axle loads and other loading information are provided in Table 4.1. Five
positions of the live load vehicle were modeled to match the experimental load positions.
The loading positions from LC 6 to LC 10 were shown in Figure 3.3: truck tandem axles
are centered on 1) left springline, 2) left shoulder, 3) crown, 4) right shoulder, and 5)
right springline.

88
6.4 Finite element analysis results
In this study, 14 of the 39 test culverts were modeled and analyzed using CANDE
program. Analysis results were compared to experimental and theoretical results. Culvert
MAD-038-0935 is selected as an example to illustrate typical deformed shapes under five
critical loading conditions. The maximum displacements and deformed shapes of the
culvert corresponding to the live load locations (LC 6 to LC 8) are shown in Figure 6.6.
The maximum deflections occur at the crown and both shoulder of the culvert. The
maximum deflection of the culvert MAD-038-0935 is 0.077 in. (1.956 mm) when the
center axle of the truck tandem is located over the crown (LC 8). The simulated
deflection is about 2.4 times larger than that of measured deflection (0.032 in. (0.813
mm)).
Measured and predicted deflections at the crown are compared in Table 6.2.
Theoretical deflections are calculated from Equation 2.1 (Spangler theory). When the
tandem axles of the truck positioned over the crown, measured deflections are smaller
than the predicted and theoretical deflections. Deflections from CANDE analyses for
both soils (SM and SC) and from Equation 2.1 are always larger than the experimental
results (Figure 6.7)
Chelliah (1992) conducted a detail parametric study using CANDE program and
reported that increase in the modulus number ( K ) in the Duncan hyperbolic model gave
better agreement between the measured and predicted deflections in the back-analysis.
However, deflections predicted by CANDE are always larger than the experimental
results even if different factors to enhance soil strength are used. Similarly, in this
research, the calculated displacements were consistently larger than those measured in the
field.
Thrust forces are calculated for the 39 test culverts using Equation 5.7 and
circumferential strains measured during field tests. The maximum experimental thrust
forces presented in Table 6.3 and Figure 6.8 are calculated at the crown when the truck
tandem was located right over the crown (LC 8 in Figure 6.6). Table 6.2 and Figure 6.8
show that, in general, experimental thrust forces are similar to predicted and theoretical
thrust forces. Chang and Selig (1980) compared CANDE predictions to measured
response of a 26 ft-span culvert and reported that thrusts calculated from CANDE agrees

89
well with the measured thrusts after the backfill is placed above the crown. The
difference between the calculated and measured thrusts is reasonable due to the fact that
CANDE and theoretical calculations are based on design values, which do not reflect
damage such as section area loss, permanent deflection, and seam connection problems.
Thrust forces for two different soils (SM and SC) are similar (Table 6.3), suggesting that
the effect of soil type on thrust forces is negligible. As supported by this conclusion,
generally, soil type is not considered in load rating methods.

90
LC 6 LC 7

Max. deflection: 0.050 in. (1.276 mm) Max. deflection: 0.066 in. (0.682 mm)

LC 8 LC 9

Max. deflection: 0.077 in. (1.956 mm) Max. deflection: 0.071 in. (1.816 mm)

Magnification = 80
LC 10

Max. deflection: 0.058 in. (1.472 mm)

Figure 6.6: Deformed shapes of MAD-38-0935 under different truck loadings.

91
Table 6.2: Measured, predicted (CANDE), and theoretical (Equation 2.1) crown
deflections.

Experiment CANDE results Theory


Culvert SM SM SC SC
number Culvert name (in.) (mm) (in.) (mm) (in.) (mm) (in.) (mm)
1 ADA-041-0096 0.011 0.279 0.056 1.412 0.138 3.495 0.201 5.105
2 ALL-309-1664 0.013 0.330 - - - - 0.19 4.826
3 ALL-696-0410 0.024 0.610 0.075 1.895 0.183 4.648 0.267 6.782
4 ATB-193-236 0.028 0.711 - - - - 0.238 6.045
5 ATB-307-1801 0.015 0.381 - - - - 0.209 5.309
6 CLI-068-0942 0.001 0.025 0.009 0.229 0.019 0.483 0.048 1.219
7 COL-170-1981 0.037 0.940 - - - - 0.381 9.677
8 COL-172-1209 0.029 0.737 0.104 2.639 0.244 6.203 0.278 7.061
9 DEF-018-1560 0.014 0.356 - - - - 0.267 6.782
10 DEL-037-0333 0.003 0.076 - - - - 0.058 1.473
11 FAY-734-0790 0.049 1.245 0.079 1.996 0.026 0.660 0.207 5.258
12 FRA-062-2688 0.02 0.508 0.046 1.176 0.112 2.840 0.187 4.75
13 GAL-218-0578 0.038 0.965 - - - - 0.198 5.029
14 HIG-134-0782 0.045 1.143 0.085 2.167 0.207 5.268 0.243 6.172
15 HOC-093-1975 0.028 0.711 0.064 1.626 0.154 3.899 0.202 5.131
16 HOC-664-1172 0.038 0.965 - - - - 0.264 6.706
17 JAC-327-0077 0.073 1.854 - - - - 0.237 6.02
18 KNO-013-2162 0.011 0.279 - - - - 0.138 3.505
19 KNO-062-1480 0.001 0.025 - - - - 0.055 1.397
20 LAK-608-0303 0.055 1.397 - - - - 0.3 7.62
21 LIC-079-2276 0.009 0.229 - - - - 0.202 5.131
22 LIC-661-0805 0.017 0.432 - - - - 0.133 3.378
23 MAD-038-0935 0.032 0.813 0.077 1.963 0.209 5.311 0.214 5.436
24 MAD-056-1490 0.009 0.229 - - - - 0.117 2.972
25 MAD-665-1071 0.027 0.686 0.049 1.242 0.127 3.226 0.159 4.039
26 MAR-229-0093 0.016 0.406 - - - - 0.175 4.445
27 MAR-309-1248 0.013 0.33 - - - - 0.176 4.47
28 MAR-423-0214 0.007 0.178 0.035 0.894 0.085 2.169 0.14 3.556
29 MAR-423-0379 0.019 0.483 - - - - 0.173 4.394
30 MRW-019-0456 0.016 0.406 - - - - 0.211 5.359
31 MRW-095-1637 0.016 0.406 - - - - 0.21 5.334
32 MRW-095-1796 0.008 0.203 0.041 1.039 0.052 1.323 0.105 2.667
33 NOB-083-0419 0.035 0.889 - - - - 0.274 6.96
34 PIC-138-0091 0.014 0.356 - - - - 0.151 3.835
35 PIC-159-0868 0.004 0.102 - - - - 0.075 1.905
36 PIC-752-0748 0.004 0.102 - - - - 0.076 1.93
37 SAN-412-0575 0.089 2.261 0.161 4.087 0.376 9.553 0.389 9.881
38 SAN-510-0412 0.047 1.194 - - - - 0.347 8.814
39 TRU-193-0215 0.019 0.483 0.057 1.455 0.137 3.467 0.21 5.334

92
Table 6.3: Measured, predicted, and theoretical thrust forces at the crown.

Experiment CANDE results Theory


Culvert Culvert name kN/ kN/ kN/
lb/in. kN/m lb/in. lb/in. lb/in.
number m m m
1 ADA-041-0096 77 13 79 14 50 8.8
2 ALL-309-1664 109 19.1 - - - - 51 8.9
3 ALL-696-0410 81 14.2 104 18 98 17 101 17.7
4 ATB-193-236 - - - - 91 15.9
5 ATB-307-1801 112 19.6 - - - - 67 11.7
6 CLI-068-0942 0.0 0 0 5 1 3 0.5
7 COL-170-1981 146 25.6 - - - - 467 81.8
8 COL-172-1209 204 35.7 125 22 124 22 134 23.5
9 DEF-018-1560 26 4.6 - - - - 79 13.8
10 DEL-037-0333 66 11.6 - - - - 12 2.1
11 FAY-734-0790 99 17.3 109 19 105 18 115 20.1
12 FRA-062-2688 127 22.2 66 12 69 12 37 6.5
13 GAL-218-0578 100 17.5 - - - - 52 9.1
14 HIG-134-0782 213 37.3 118 21 110 19 136 23.8
15 HOC-093-1975 128 22.4 77 13 82 14 57 10.0
16 HOC-664-1172 191 33.4 - - - - 114 20.0
17 JAC-327-0077 277 48.5 - - - - 113 19.8
18 KNO-013-2162 101 17.7 - - - - 25 4.4
19 KNO-062-1480 - - - - 3 0.5
20 LAK-608-0303 - - - - 166 29.1
21 LIC-079-2276 - - - - 50 8.8
22 LIC-661-0805 162 28.4 - - - - 41 7.2
23 MAD-038-0935 146 25.6 100 18 92 16 97 17.0
24 MAD-056-1490 - - - - 22 3.9
25 MAD-665-1071 70 12.3 72 13 71 12 48 8.4
26 MAR-229-0093 119 20.8 - - - - 75 13.1
27 MAR-309-1248 49 8.6 - - - - 59 10.3
28 MAR-423-0214 48 8.4 50 9 53 9 28 4.9
29 MAR-423-0379 97 17.0 - - - - 59 10.3
30 MRW-019-0456 81 14.2 - - - - 69 12.1
31 MRW-095-1637 100 17.5 - - - - 93 16.3
32 MRW-095-1796 50 8.8 49 9 53 9 14 2.5
33 NOB-083-0419 247 43.3 - - - - 240 42.0
34 PIC-138-0091 79 13.8 - - - - 50 8.8
35 PIC-159-0868 - - - - 9 1.6
36 PIC-752-0748 34 6.0 - - - - 14 2.5
37 SAN-412-0575 175 30.6 184 32 165 29 244 42.7
38 SAN-510-0412 95 16.6 - - - - 237 41.5
39 TRU-193-0215 182 31.9 69 12 75 13 53 9.3

93
0.40 10
CANDE (SM soil)
0.35 CANDE (SC soil)

Deflection at the crown (mm)


8
Deflection at the crown (in.)
0.30 Experiment
Theory
0.25
6
0.20

0.15 4

0.10
2
0.05

0 0
3 8 12 15 25 32 39
Culvert number
Figure 6.7: Maximum deflection on the culvert crown obtained from experimental tests
and numerical analyses.

300
50
CANDE (SM soil)
250 CANDE (SC soil)

Deflection at the crown (mm)


Deflection at the crown (in.)

Experiment 40
200 Theory

30
150

20
100

50 10

0 0
0 5 10 15 20 25 30 35 40
Culvert number
Figure 6.8: Thrust forces acting on the culvert crown obtained from experimental tests
and numerical analyses.

94
CHAPTER 7

ODOT’s LOAD RATING PROCEDURE

7.1 Introduction
Load rating for the design and evaluation of a culvert can be defined as the ratio
of actual live load carrying capacity of a culvert to the factored design live and impact
load. The load rating is generally expressed as a Rating Factor (RF), which can be
considered similar to a factor of safety that relates the existing load capacity to external
load demand (Akgül and Frangopol, 2004). For example, if a standard HS-20 truck load
is used for design of a culvert, and if the predicted RF is 5, theoretically, the culvert can
support a live load five times the design load for HS-20 truck.
According to the AASHTO Manual for Condition Evaluation of Bridges (1994),
every bridge should be rated at inventory and operating load levels. The inventory load is
defined as a live load that can safely cross over the culvert for an indefinite period. The
operating load is the maximum permissible live load that the culvert can support.
Allowing an unlimited number of vehicles to use the bridge at the operating level may
shorten the life of the bridge (AASHTO, 1994).
There are three methods available for design of bridges: 1) allowable stress design
(ASD) or working stress design, 2) load factored design (LFD), and 3) load and
resistance factored design (LRFD). On the other hand, only one load rating method has
been developed for culverts, which is based on LFD method. The National Corrugated
Steel Pipe Association (NCSPA, 1995) reported evaluation methods for load rating based
on AASHTO Standard Specifications for Highway Bridges (1992) and available research
results (Duncan and Drawsky, 1983; Cowherd and Degler, 1986). According to the
NCSPA design procedure, RF is the lower of the load rating factor based on wall strength
(RFW) and the load rating factor based on minimum cover (RFC). RFW is calculated
from culvert wall thrust force resulting from factored dead and live loads, and RFC is

95
determined from the cover depth over the culvert. The details of the load rating
procedures proposed or used by NCSPA, AASHTO, and Ohio Department of
Transportation (ODOT) are discussed in the next sections.

7.2 ODOT’s CMP-Excel program


The Ohio Department of Transportation uses a load rating procedure to evaluate
more than 1000 culverts in its inventory. Corrugated metal pipe (CMP) culverts are rated
using an in-house computer program called CMP-Excel. The program can calculate RFs
for most flexible long-span culverts including circular pipes, structural plate culverts, and
spiral rib culverts. The CMP-Excel program consists of three parts. The first part includes
basic information on material properties, culvert geometry and structural design
parameters calculated during the initial culvert design or revised during the latest
inspection if the culvert condition has changed since its construction. The information
includes type of culvert, span length, longitudinal length of the structure, clear cover
above the crown (soil backfill height plus pavement thickness), the measured top radius,
and the measured vertical deflection of the crown. The second part of CMP-Excel
includes load rating factor calculations using the current properties of the culvert. The
thrust forces required for RF calculations are based on design dead and live loads from
AASHTO Standard Specifications for Highway Bridges (2002). For new culverts, initial
design parameters are used. For existing culverts, the properties measured during the
latest annual culvert inspection are used. For example, if the culvert wall thickness is
reduced due to corrosion, culvert wall cross sectional area, and in turn, wall strength and
buckling strength are reduced. The last part of CMP-Excel calculates the inventory and
operating loads for the culvert. Figure 7.1 summarizes ODOT’s current culvert load
rating procedure. The parameters shown in the figure are described and discussed in the
following sections. The next four sections describe the CMP-Excel program and its
components.

96
Input basic geometrical and material properties
Follow the bridge inspection (clear cover, span length, longitudinal length,
report (BR-86) and decide what structure type etc.) (See Section 7.3)
properties need to be changed
based on engineering judgment. Check BR-86, and select a new gauge number, or
change the gauge number if there is a change in
culvert condition (Section 7.4)
Check if there is a change in
geometry
Calculate the seam strength, φseam S s
(Section 7.4.1)

Calculate the wall yield strength, φwall f y A


(Section 7.4.2)

Calculate the buckling stress, f b


(Section 7.4.3)

No
Any increase
in measured Calculate the wall buckling strength, φbkl fb A
deflection? (Section 7.4.3)

Yes Thrust capacity, Tcap is the lower of


Calculate a new buckling factor ① Seam strength ( φseam S s )
φbkl = 0.95 − 5.6δ ② Wall yield strength ( φwall f y A )
(Sections 7.3 and 7.4.3) ③ Buckling strength ( φbkl fb A )
(Section 7.5.1)

Calculate the wall thrust due to earth cover, TE


Calculate the wall thrust due to live load, T( L +i )
(Sections 7.5.2 and 7.5.3)

Determine operating load rating based on wall


strength (RFW) and minimum cover (RFC)
(Section 7.6.1)

Determine inventory load ratings based on wall


strength (RFW) and minimum cover (RFC)
(Section 7.6.2)

Figure 7.1: ODOT load rating flow chart.

97
7.3 Basic information needed for load rating evaluation
User of the CMP-Excel needs to enter the information obtained from the culvert
design or annual bridge inspection report (BR-86). The required input data includes type
of culvert, longitudinal length of the culvert ( Lc ), clear cover above the culvert crown

( H ), and span length ( Sc ) as shown in Figure 7.2. The CMP-Excel provides various load
rating methods for different structures, such as long span structural plate culverts,
structural plate culverts, round pipes, and spiral rib aluminum and steel pipes. This study
focuses on load rating of round pipes and structural plate culverts only. When a culvert is
already in-service, the actual culvert information obtained during the annual inspection
has to be used. Generally, H , Sc , and Lc are not measured in the field during inspections

after the initial construction. However, the culvert height H c may be measured if signs of
deflection, flattening, or buckling are found in the culvert.

Figure 7.2: Notation relating to culvert geometry.

If the culvert experiences permanent deformations during or after construction, a new


load rating is assigned based on the changes in top radius ( Rt ) and the culvert deflection

rate reported in BR-86. These changes affect the buckling factor ( φbkl ) and the span

length ( Sc ). φbkl is calculated from the following equation for structural plate culverts or

98
from Figure 7.3 for round pipes
φbkl = 9.5 − 5.6δ (7.1)

where δ is the culvert deflection rate, δ = Δ / H c , where Δ is the vertical culvert

deflection measured at the crown, and H c is the culvert height (Figure 7.2). Ideally, the

current initial measured H c should be used, however H c specified in the original design

is typically used. Buckling strength of the culvert and φbkl will be discussed further in
Section 7.4.3. According to NCSPA (1995) Design Data Sheet No. 19 I.A.2.,
asymmetrical structures, structures deflected more than 5% from the original condition,
or those that show localized distortions require that the actual maximum top radius,
Rt ,actual is determined in those distorted areas as shown in Figure 7.4. NCSPA

recommends that, for a conservative evaluation, the designer use two times the actual
maximum radius rather than the span length in structural design. However, it is not easy
to determine the actual radius of culvert from field measurement. Thus, most culverts are
evaluated using two times the design radius when the culvert is deflected or buckled.

1
Buckling strength reduction factor

0.8

0.6

0.4

0.2

0
0 2 4 6 8 10
Culvert deflection (%)

Figure 7.3: Reduction factor for buckling strength ( φbkl ) due to deflection in round pipes
(NCSPA, 1995).

99
m l2
Rt ,actual = +
2 8m
l = length of straight edge
m = mid ordinate

Figure 7.4: Determination of actual culvert radius by field measurement (NCSPA, 1995).

7.4 Physical strength calculations


In the second part of CMP-Excel, the physical strength of culvert is calculated
considering wall strength, seam strength, and buckling strength following the AASHTO
(2002) design requirements. Thrust forces due to design live load and dead load are also
determined. If the culvert was not previously rated, the material properties or gauge
number of the plate selected during the initial design are used. If the culvert needs to be
re-rated because of observed damage such as steel corrosion, pitting, bolting problems,
gaps between plate connections, holes in the plate, etc., a new gauge number is selected
by the inspector or engineer who evaluate the current culvert condition based on
engineering judgment. The main focus of this study is on structural plate culverts, that is,
bolted structures with 6 × 2 in. (152 × 51 mm) corrugation as shown in Figure 7.5. The
sectional properties of arc-and-tangent type of corrugation are given in Table 7.1. The
properties include area, A , moment of inertia, I , section modulus, S , and radius of
gyration, r (= I A ) . In this research, it is assumed that the gauge number has already
been decided from initial design calculations or from the latest field inspection. If the
annual inspection indicates some form of damage in the culvert, the thickness of the plate
or gauge number in Table 7.1 is reduced based on engineering judgment. The physical
strength and thrust forces in the culvert are calculated and the culvert is evaluated using
the revised gauge number as described in the next sections.

100
Table 7.1: Sectional properties of 6 × 2 in. (152 × 51 mm) corrugation plate (AISI, 1994,
Table 2.10).

Gauge Specified Uncoated Area of Moment of Section Radius of


number thickness thickness section inertia modulus gyration
(in.) (in.) (in.2/ft) (in.4/ft) (in.3/ft) (in.)
12 0.111 0.1046 1.556 0.725 0.689 0.682
10 0.140 0.1345 2.003 0.938 0.879 0.684
8 0.170 0.1644 2.449 1.154 1.066 0.686
7 0.188 0.1838 2.739 1.296 1.187 0.688
5 0.218 0.2145 3.199 1.523 1.376 0.690
3 0.249 0.2451 3.658 1.754 1.562 0.692
1 0.280 0.2758 4.119 1.990 1.749 0.695
0.310 0.3125 4.671 2.280 1.968 0.698
0.380 0.3750 5.613 2.784 2.340 0.704

Gauge Specified Uncoated Area of Moment of Section Radius of


number thickness thickness section inertia modulus gyration
(mm) (mm) (cm2/m) (cm4/m) (cm3/m) (mm)
12 2.819 2.657 32.9 99.0 37.0 17.323
10 3.556 3.416 42.4 128.1 47.3 17.374
8 4.318 4.176 51.8 157.6 57.3 17.424
7 4.775 4.669 58.0 177.0 63.8 17.475
5 5.537 5.448 67.7 208.0 74.0 17.526
3 6.325 6.226 77.4 239.5 84.0 17.577
1 7.112 7.005 87.2 271.8 94.0 17.653
7.874 7.938 98.9 311.4 105.8 17.729
9.652 9.525 118.8 380.2 125.8 17.882

Figure 7.5: Geometry of 6 × 2 in. (152 × 51 mm) corrugation plate.

101
7.4.1 Seam strength
According to Section 12.3.3 of AASHTO Specifications (2002), the longitudinal
seams are required to have seam strength, S s to satisfy Equation 7.2.

φseam S s ≥ Tl (7.2)

where φseam is equal to 0.67 for CMP culverts. For each gauge number, the values of S s
are given in Table 7.2. AASHTO Specifications do not provide any detail or guidance for
the calculation of Tl , which is wall thrust due to factored external dead and live loads.
The simplest method uses ring compression theory which is discussed in Section 2.3.3
and yields Tl from the following equation

Sc
Tl = [γ L (1 + i) pL + γ D pD ] (7.3)
2
where Sc is the span length, γ L is the load factor for live load, γ D is the load factor for

dead load, i is the impact factor, pL and pD are equivalent uniformly distributed

pressures at the crown due to live load and dead load, respectively. Calculation of pL and

pD will be discussed in Sections 7.5.3 and 7.5.2, respectively.

7.4.2 Wall strength


Wall strength is calculated by the following equation (AASHTO, Section 12.3.1,
2002).
φwall f y A ≥ Tl (7.4)

where φwall is the capacity modification factor for wall, and is equal to 1.0. A is the

cross-sectional area of the culvert wall per unit length, and f y is the specified minimum

yield strength of the wall plate. Tl can be calculated from Equation 7.3.

102
Table 7.2: Minimum longitudinal seam strengths for 6 × 2 in. (152 × 51 mm) corrugation
plate (Section 12.6.2 of AASHTO, 2002).

Gauge Specified Bolt


4 bolts/foot 6 bolts/foot 8 bolts/foot
number thickness diameter
(kips/ft) (kips/ft) (kips/ft)
(in.) (in.)
12 0.111 3/4 43.0
10 0.140 3/4 62.0
8 0.170 3/4 81.0
7 0.188 3/4 93.0
5 0.218 3/4 112.0
3 0.249 3/4 132.0
1 0.280 3/4 144.0 180.0 194.0
0.318 7/8 235.0
0.380 7/8 285.0

Gauge Specified Bolt


4 bolts/foot 6 bolts/foot 8 bolts/foot
number thickness diameter
(kN/m) (kN/m) (kN/m)
(mm) (mm)
12 2.819 19.050 627.5
10 3.556 19.050 904.8
8 4.318 19.050 1182.1
7 4.775 19.050 1357.2
5 5.537 19.050 1634.5
3 6.325 19.050 1926.4
1 7.112 19.050 2101.5 2626.9 2831.2
8.077 22.225 3429.6
9.652 22.225 4159.3

7.4.3 Buckling strength


The buckling strength is calculated from Section 12.3.2 of AASHTO (2002).
φbkl fb A ≥ Tl (7.5)

where φbkl is the capacity modification factor for buckling, and fb is the buckling stress
given by the following equations.
f u2 r 24 E
( K s Sc / r )
2
fb = fu − if Sc < (7.6)
48E Ks fu

12 E r 24 E
fb = if Sc ≥ (7.7)
( K s Sc / r )
2
Ks fu

103
where Sc is span length, K s is the soil stiffness factor (recommended to use 0.22 for
culverts by AASHTO (Section 12.3.2, 2002)), E is the modulus of elasticity of wall plate,
r is the radius of gyration of corrugation, and fu is the ultimate tensile strength.

If an existing culvert does not have any buckling damage, φbkl is set equal to 1.0.

If there is flattening, deflection or localized global buckling of the culvert, φbkl is


calculated from Equation 7.1. According to CMP-Excel program, the maximum
allowable deflection ratio δ is 0.1 for load rating. If the deflection ratio of the culvert is
higher than 10%, the culvert may need to be closed, repaired or replaced.

7.5 Thrust forces

7.5.1 Thrust capacity of wall


Three different physical strengths are calculated for each culvert as presented in
Section 7.4. According to NCSPA (1995), the capacity modification factor for wall and
buckling strengths are not included in the calculation for wall thrust capacity. However,
the modification factors are generally considered for the load rating calculations. In this
study the factors are included in thrust capacity calculation because this procedure is used
in the CMP-Excel program. The smallest of three strengths controls the maximum
allowable thrust capacity for culvert, Tcap (Equation 7.8).

⎧ Wall yield strength = φwall f y A



Tcap = minimum of ⎨Buckling strength = φbkl fb A (7.8)
⎪Seam strength =φ S
⎩ seam s

7.5.2 Wall thrust due to dead load


The load rating evaluation requires the calculation of thrust force created by the
vertical earth pressure over the culvert as a prism load. The prism load is the weight of
the rectangular prism of soil above the culvert as shown in Figure 7.6. The pipe wall
thrust due to earth cover, TE is

Sc
TE = (γ s H + LL ) (7.9)
2

104
where γ s is the soil density, H is the clear cover above the crown, and LL is live load
surcharge.

Figure 7.6: Prism load and live load surcharge over culvert.

7.5.3 Wall thrust due to live load plus impact


In order to check the pressure over the crown due to external live load, the
equivalent live-load pressure, pL , needs to be calculated by assuming that a wheel load is
uniformly distributed over a rectangular area. According to AASHTO design standard
(2002), each side of the loading area is increased by 1.75 times H as shown in Figure 7.7.
When such areas due to several wheel loads overlap, the total load is assumed to be
distributed uniformly over the area. The depth at which neighboring wheels participate in
pressure is defined as the overlapping depth, H ol , and is calculated from the following
equation.

H ol =
( 0.5ws − ww ) (7.10)
0.5φs

where ws is wheel spacing on axle, ww is one wheel contact width in the axle direction,

and φs is spreading factor. For a standard AASHTO HS-20 truck load, ws = 6 ft (1.83 m),

105
ww = 10 in. (254 mm), and φs = 1.75. Thus, H ol = 2.476 ft (755 mm).

Live load impact factor, i , is specified to be zero when H is greater than 3 ft


(0.91 m). For H less than 3 ft (0.91 m), the impact factor varies depending on fill height
over the crown. For each 1 ft (0.3 m) decrease in H , the impact factor increases by 0.1.
Thus, i is 0.3 when H is zero.
The pressure due to HS-20 truck load plus impact, pL +i can be obtained from the
following equation.
0.5PHA ⋅ i
pL + i = if H < H ol (7.11)
(10 /12 + 2 × 0.875 H )(20 /12 + 2 × 0.875 H )
PHA ⋅ i
pL + i =
2 (10 /12 + 1.75 H )( 20 /12 + 1.75 H ) − (10 /12 + 1.75H )(1.75( H − H ol ) )

if H ≥ H ol (7.12)

where PHA is heavy axle load (32,000 lb (142 kN) for HS-20).

The culvert wall thrust due to live load plus impact, TL +i can be calculated by the
ring compression theory which was described in Section 2.3.3.
1
TL +i = pL + i S c (7.13)
2

7.6 Rating factors for ring compression structures


Evaluation of the load capacity of existing culverts is very important for ODOT
engineers to manage a large number of in-service culverts in Ohio. Understanding the
true load carrying capacity of these structures may ensure the safety of the travelling
public. In this section, methods for calculation of the load carrying capacity and rating
factor will be discussed.
The load rating is generally expressed as a rating factor for a particular live load
model, using the general load-rating equation as shown in the following equation
(AASHTO, 1994),
Cm − β D Dm
RF = (7.14)
β L Lm (1 + i )

106
where Cm is the capacity of the structure, Dm is the dead load on structure, Lm is the live

load on the structure, i is the impact factor to be used with the live load effect, β D is the

factor for dead loads, and β L is the factor for live load.

(a) Plan view

(b) Section A-A (c) Section B-B

Figure 7.7: HS-20 load distribution through soil cover over culvert in longitudinal
direction.

107
7.6.1 Operating load rating
In this section, RFs for operating load rating are introduced based on the NCSPA
design procedure (1995). According to NCSPA, the controlling RF of the load rating is
the smaller of two different RF values; one based on wall strength (RFW) and the other
based on minimum cover requirements (RFC). RFW (NCSPA III. B.1.a) is calculated as
Tcap − 1.95TE
RFW = (7.15)
1.3TL +i
The load factors in Equation 7.15 can be explained using AASHTO specifications (2002).
According to AASHTO, the combined load factor for earth pressure is γβ E = 1.95 where

γ = 1.3 for load factor design, and β E is 1.67 for culverts. Load factor for live load is
γβ L +i = 1.3 where β L +i = 1.0 for service load.
RFC (NCSPA III. B.1.b) is determined as
H2
RFC = (7.16)
C ⋅ h2
where h is the AASHTO minimum cover level for the structure with structural plate pipe
and shall be Sc / 8 but not less than 12 in. (0.3 m) (AASHTO 12.6.1.5, 2002).

H
C = 2.36 + 0.528 ≤ 1.0 . (7.17)
Sc
According to NCSPA Design Data Sheet No. 19, C is calculated based on the plastic
moment strength suggested by Duncan and Drawsky (1983). Although the plastic
moment is not directly used for the load rating evaluation, it is described here to help
understand how Equation 7.16 is derived.
The required plastic moment capacity of the culvert M p is calculated as
2
⎛S ⎞
M p = K3 ⎜ c ⎟ (7.18)
⎝H⎠
Fp PHA d
where K 3 = , Fp is factor of safety against development of a plastic hinge (see
c
Figure 7.8), PHA is heavy axle load (32,000 lb (142 kN) for HS-20), d is corrugation

depth (2 in. (51 mm) for 6 × 2 in. (152 × 51 mm) corrugation plate), c is a coefficient
with a unit of length whose value depends on the degree of compaction of the backfill.

108
For example, c = 69 ft (21.03 m) for relative compaction (RC) of 90 percent standard
Proctor, and c = 115 ft (35.05 m) for RC of 95 percent standard Proctor. The maximum
allowable plastic moments for standard corrugated plates (gauge number) are shown in
Table 7.3 (Duncan and Drawsky, 1983).
The required plastic moment capacity for the AASHTO minimum cover is
Fp PHA d ⎛ Sc ⎞ 2
M p ,c = ⎜ ⎟ from Equation 7.18. The design M p ,d for the design cover depth
c ⎝ h⎠
( H ) must smaller than M p ,c to satisfy structural safety. Therefore, RF determined from

the required moment capacity can be expressed by


2
Fp PHA d ⎛ Sc ⎞
M p ,c c ⎜⎝ h ⎟
⎠ =H
2
RF = = 2
(7.19)
M p ,d Fp PHA d ⎛ Sc ⎞ h2
c ⎜⎝ H ⎟

This RF is determined based on cover depth and does not consider the capacity
modification factor. Thus, Equation 7.19 is the inventory rating factor based on cover,
and needs to be rewritten as the following equation.
H2
RFCi = 2 (7.20)
h
To reduce the factor of safety for operating loads, a capacity modification factor
needs to be introduced in Equation 7.20. According to AASHTO(2002), the minimum
H
cover limits, h are Sc / 8 . The cover ratio ( ) for the minimum cover is
Sc
Sc / 8 1
= = 0.125 . From Figure 7.8, the required factor of safety, Fp for AASHTO
Sc 8

minimum cover is Fp ,i = 1.488 for inventory loads (Figure 7.8b) and Fp ,o = 1.225 for

operating loads (Figure 7.8a). The balance factor for the AASHTO minimum cover, C ,
Fp ,o 1.225
which is the ratio of Fp ,o and Fp ,i is = = 0.823 when the cover ratio is 0.125.
Fp ,i 1.488

The balance factor is 1.0 when the cover ratio is 0.2. Thus, the balance factor, C can be
derived by using a linear relationship as shown in Figure 7.9, and is already shown in

109
Equation 7.17. Thus, RF for operating loads based on cover shown in Equation 7.16 can
be derived.

Table 7.3: Maximum allowable moment strength (Duncan and Drawsky, 1983).

Gauge number 12 10 8 7 5 3 1
(in.) 0.111 0.140 0.170 0.188 0.218 0.249 0.280
Plate thickness
(mm) 2.819 3.556 4.318 4.775 5.537 6.325 7.112
(k-ft/ft) 2.66 3.44 4.22 4.73 5.54 6.36 7.18
Mp
(kN-m/m) 11.83 15.30 18.77 21.04 24.64 28.29 31.94

7.6.2 Inventory load rating


The RF for inventory loading can be determined from the lower value of
RFW and RFC . According to AASHTO Specification (2002), the factor for live load
5
plus impact for a culvert is 1.67 or for design (evaluation) condition (AASHTO
3
3.22.1). The thrust force due to live load plus impact is then increased by this factor. Thus,
inventory rating factor based on wall strength is
Tcap − 1.95TE 3
RFWi = = RFW (7.21)
1.67 × 1.3TL +i 5
The inventory rating factor based on minimum cover requirements was explained in
Section 7.6.1, and is given in Equation 7.20.

7.7 Discussion of ODOT’s load rating method.


Various parameters and the source and implications of each equation used for
load rating evaluation were discussed in previous sections. If the parameters determined
from original culvert design are used for the load rating procedures for the first time, the
application of procedure is straightforward (Figure 7.1). After a culvert is constructed and
in service for some time, however, the culvert could be damaged due to traffic loading,
water pressure, soil pressure, steel corrosion, and other loading and environmental
conditions. Thus, the condition of the culvert deteriorates as it gets older. If any culvert

110
(a) Required values of Fp for operating loads with axle loads up to 32 kips.

(b) Required values of Fp for inventory loads with axle loads in excess of 32 kips.

Figure 7.8 Required values of Fp .

111
Figure 7.9: Balance factor C .

damage is reported during annual inspection, an ODOT engineer may decide to change
the load rating. ODOT then revises the load rating of the culvert by changing the span
length, top radius of the culvert, deflection rate and/or gauge number based on
engineering judgment. Although some culverts may have the same general appraisal
number and similar corrosion damage, area reduction rate, and load rating may vary
depending on the engineer who rates the culvert. The next section investigates how
engineering judgment affects the load rating evaluation of in-service culverts.

7.7.1 Change in gauge number


ODOT inspects culverts in its inventory annually, and records inspection results
in a BR-86 bridge inspection report. If an inspected culvert has damage, its load rating
needs to be checked. Before deciding whether the gauge number needs to be changed, the
engineer checks the culvert deflection, corrosion damage, seam condition, etc., reported
in the BR-86 report. However, most of the decisions are based on engineering judgment.
Although some culverts may have the same general appraisal number (GAN) and similar
corrosion damage, change in the gauge number, hence the area reduction rate, may vary
depending on the engineer who rates the culvert. Table 7.4 shows design gauge numbers
and current gauge numbers used for the load rating evaluation of 39 culverts tested in this
research. The design span lengths and culvert heights shown in the table were obtained
from culvert design plans unless noted otherwise. The age and other properties of the
culverts were provided in Chapter 4. As an example, culverts ATB-193-2236 and ATB-

112
307-1801 have the same GAN of 4 (Table 7.4). Both culverts have heavy pack rust and
perforations. The gauge number for ATB-193-2236 was increased from 3 to 5 to reflect
this damage and potential reduction in metal cross sectional area. However, the gauge
number stayed the same for ATB-307-1801. If the same engineer rated both culverts, a
similar change in gauge number would be expected for both culverts. Thus, a better
guideline needs to be provided for the engineer to reduce bias or mistakes and to provide
consistency for load rating evaluation.

7.7.2 Capacity of wall strength, Tcap

The capacity of wall strength is determined by the lower of wall yield strength
( = φwall f y A , Section 7.4.2), buckling strength ( = φbkl f b A , Section 7.4.3), and seam

strength ( = φseam S s , Section 7.4.1). According to ODOT’s current load rating method,

parameters φwall = 1.0 and φseam = 0.67 stay the same value although the culvert
condition gets worse. The other parameters φbkl , A , and S s can be possibly changed as a
culvert ages. Our evaluation of 39 culverts shows that seam strength is almost always
smaller than wall yield strength and buckling strength as shown in Table 7.5. Thus, seam
strength appears to generally control wall strength capacity. When the average strength
values for design are compared, the wall yield strength and buckling strength are 45%
and 78% higher than the seam strength, respectively. According to ODOT’s current load
rating procedure, values of wall yield strength are generally much higher than the
buckling and seam strength (Table 7.5). Thus, wall yield strength never controls wall
strength capacity for the 39 culverts considered in this study. The engineer may not even
need to calculate the wall yield strength.
When the culvert has local buckling, flattening, or crown deflection, the capacity
modification factor for buckling strength, φbkl , has to be changed based on Equation 7.1.
Even if the culvert has some deflection or local buckling, the buckling strength is not
likely to control wall strength capacity because the buckling strength is still much higher
than the seam strength. However, the buckling strength may control the load rating
evaluation if culvert deflection ratio is higher than 7% ( φbkl = 0.558 ). Therefore, only
very large deflections appear to affect wall strength capacity.

113
Table 7.4: Comparison of initial design and current gauge numbers.

Design information Current information


Span Culvert Gauge GAN Span Deflection Gauge
Culvert name
Height Number Number
(ft) (ft) (ft) (%)
ADA-041-0096 11.83* 7.58* 8* 4 12.45 5 10
* *
ALL-309-1664 16.58 10.08 5* 5 16.66 0 7
ALL-696-0410 12.50* 7.92* 7* 5 12.58 0 8
* *
ATB-193-2336 15.83 9.83 3* 4 15.92 0 5
ATB-307-1801 16.58 10.08 5 4 16.58 0 5
CLI-068-0942 11.25* 11.25* 3* 7 11.81 5 5
COL-170-1981 13.42 8.42 5 5 14.13 5 7
COL-172-1209 16.58 10.08 5 5 17.49 5 7
DEF-018-1560 10.67 6.92 12 5 11.20 5 12
DEL-037-0333 21.50 7.75 7 5 28.50 0 7
FAY-734-0790 12.50 7.92 7 5 13.21 5 8
FRA-062-2688 12.50 7.92 8 5 13.21 5 10
GAL-218-0578 15.50 9.42 3 5 15.50 0
HIG-134-0782 12.50 7.92 7 5 13.21 5 8
HOC-093-1975 14.25 8.92 7 3 - - -
HOC-664-1172 14.25 8.92 7 5 14.32 0 8
JAC-327-0077 13.42 8.42 5 5 14.13 5 7
KNO-013-2162 14.25** 8.92** 7* 6 14.96 5 8
KNO-062-1480 12.00* 12.00* 5* 6 12.60 5 7
LAK-608-0303 14.00 7.25 5 4 14.76 5.4 7
LIC-079-2276 10.25 6.75 12 6 10.26 0 12
LIC-661-0805 19.67 12.67 8 6 19.67 0 8
MAD-038-0935 10.67 6.92 8 3 11.38 5 10
MAD-056-1490 14.25 8.92 7 6 15.04 5 8
MAD-665-1071 12.42 7.92 8 6 12.58 0 8
MAR-229-0093 16.58 10.08 5 5 - - -
MAR-309-1248 10.67 6.92 8 7 10.82 0 8
MAR-423-0214 12.50* 7.92* 8* 6 12.67 0 10
MAR-423-0379 16.83* 9.8** 8* 7 16.83 0 8
* *
MRW-019-0456 12.50 7.92 10* 9 12.58 0 10
MRW-095-1637 10.67 6.92 8 7 10.67 0 8
MRW-095-1796 12.50 7.92 8 7 12.58 0 8
* *
NOB-083-0419 17.25 8.17 7* 4 17.25 0 8
PIC-138-0091 13.42* 8.42* 12* 2 14.81 10 12
* *
PIC-159-0868 12.00 12.00 7* 8 12.00 0 7
PIC-752-0748 17.42 11.50 8 8 17.42 0 8
** **
SAN-412-0575 12.83 8.33 8** 5 - - -
SAN-510-0412 11.58 7.42 5 5 12.52 7 5
TRU-193-0215 15.33** 9.25** 5** 3 - - -
*
This value is obtained from ODOT CMP-Excel sheet.
**
This value is measured by the researchers in the field.

114
Table 7.4 continued

Design information Current information


Span Culvert Gauge GAN Span Deflection Gauge
Culvert name
Height Number Number
(m) (m) (m) (%)
ADA-041-0096 3.606* 2.310* 8* 4 3.795 5 10
* *
ALL-309-1664 5.054 3.072 5* 5 5.078 0 7
ALL-696-0410 3.810* 2.414* 7* 5 3.834 0 8
* *
ATB-193-2336 4.825 2.996 3* 4 4.852 0 5
ATB-307-1801 5.054 3.072 5 4 5.054 0 5
CLI-068-0942 3.429* 3.429* 3* 7 3.600 5 5
COL-170-1981 4.090 2.566 5 5 4.307 5 7
COL-172-1209 5.054 3.072 5 5 5.331 5 7
DEF-018-1560 3.252 2.109 12 5 3.414 5 12
DEL-037-0333 6.553 2.362 7 5 8.687 0 7
FAY-734-0790 3.810 2.414 7 5 4.026 5 8
FRA-062-2688 3.810 2.414 8 5 4.026 5 10
GAL-218-0578 4.724 2.871 3 5 4.724 0
HIG-134-0782 3.810 2.414 7 5 4.026 5 8
HOC-093-1975 4.343 2.719 7 3 - - -
HOC-664-1172 4.343 2.719 7 5 4.365 0 8
JAC-327-0077 4.090 2.566 5 5 4.307 5 7
KNO-013-2162 4.343** 2.719** 7* 6 4.560 5 8
* *
KNO-062-1480 3.658 3.658 5* 6 3.840 5 7
LAK-608-0303 4.267 2.210 5 4 4.499 5.4 7
LIC-079-2276 3.124 2.057 12 6 3.127 0 12
LIC-661-0805 5.995 3.862 8 6 5.995 0 8
MAD-038-0935 3.252 2.109 8 3 3.469 5 10
MAD-056-1490 4.343 2.719 7 6 4.584 5 8
MAD-665-1071 3.786 2.414 8 6 3.834 0 8
MAR-229-0093 5.054 3.072 5 5 - - -
MAR-309-1248 3.252 2.109 8 7 3.298 0 8
MAR-423-0214 3.810* 2.414* 8* 6 3.862 0 10
MAR-423-0379 5.130* 2.987** 8* 7 5.130 0 8
* *
MRW-019-0456 3.810 2.414 10* 9 3.834 0 10
MRW-095-1637 3.252 2.109 8 7 3.252 0 8
MRW-095-1796 3.810 2.414 8 7 3.834 0 8
* *
NOB-083-0419 5.258 2.490 7* 4 5.258 0 8
PIC-138-0091 4.090* 2.566* 12* 2 4.514 10 12
* *
PIC-159-0868 3.658 3.658 7* 8 3.658 0 7
PIC-752-0748 5.310 3.505 8 8 5.310 0 8
** **
SAN-412-0575 3.911 2.539 8** 5 - - -
SAN-510-0412 3.530 2.262 5 5 3.816 7 5
TRU-193-0215 4.673** 2.819** 5** 3 - - -
*
This value is obtained from ODOT CMP-Excel sheet.
**
This value is measured by the researchers in the field.

115
Table 7.5: Design and current wall strength capacity for 39 culverts.

Design (k/ft) Current (k/ft)


Culvert name
Yield Buckling Seam Yield Buckling Seam
ADA-041-0096 80.8 102.8 54.3 66.1 55.9 41.5
ALL-309-1664 105.6 125.2 75.0 90.4 107.0 62.3
ALL-696-0410 90.4 114.1 62.3 80.8 101.9 54.3
ATB-193-2336 120.5 144.9 88.4 105.6 126.7 75.0
ATB-307-1801 105.6 125.2 75.0 105.6 125.2 75.0
CLI-068-0942 120.5 154.5 88.4 105.6 90.1 75.0
COL-170-1981 105.6 131.7 75.0 90.4 74.7 62.3
COL-172-1209 105.6 125.2 75.0 90.4 70.6 62.3
DEF-018-1560 51.3 66.2 28.8 51.3 44.1 28.8
DEL-037-0333 90.4 75.6 62.3 90.4 75.6 62.3
FAY-734-0790 90.4 114.1 62.3 80.8 67.7 54.3
FRA-062-2688 80.8 102.0 54.3 66.1 55.3 41.5
GAL-218-0578 120.5 145.7 88.4 105.6 127.6 75.0
HIG-134-0782 90.4 114.1 62.3 80.8 67.7 54.3
HOC-093-1975 90.4 111.3 62.3 - - -
HOC-664-1172 90.4 111.3 62.3 80.8 99.4 54.3
JAC-327-0077 105.6 131.7 75.0 90.4 74.7 62.3
KNO-013-2162 90.4 111.3 62.3 80.8 65.9 54.3
KNO-062-1480 105.6 134.1 75.0 90.4 76.3 62.3
LAK-608-0303 105.6 130.6 75.0 90.4 71.3 62.3
LIC-079-2276 51.3 66.5 28.8 51.3 66.4 28.8
LIC-661-0805 80.8 89.8 54.3 80.8 89.8 54.3
MAD-038-0935 80.8 104.2 54.3 66.1 56.6 41.5
MAD-056-1490 90.4 111.3 62.3 80.8 65.8 54.3
MAD-665-1071 80.8 102.0 54.3 80.8 101.9 54.3
MAR-229-0093 105.6 125.2 75.0 - - -
MAR-309-1248 80.8 104.2 54.3 80.8 104.2 54.3
MAR-423-0214 80.8 102.0 54.3 66.1 83.2 41.5
MAR-423-0379 80.8 95.3 54.3 80.8 95.3 54.3
MRW-019-0456 66.1 83.4 41.5 66.1 83.4 41.5
MRW-095-1637 80.8 104.2 54.3 80.8 104.2 54.3
MRW-095-1796 80.8 102.0 54.3 80.8 101.9 54.3
NOB-083-0419 90.4 105.8 62.8 80.8 94.5 54.3
PIC-138-0091 51.3 63.9 28.8 30.6 11.4* 17.1
PIC-159-0868 90.4 114.8 62.3 90.4 114.8 62.3
PIC-752-0748 80.8 94.2 54.3 80.8 94.2 54.3
SAN-412-0575 80.8 101.5 54.3 - - -
SAN-510-0412 105.6 134.8 75.0 105.6 74.4* 75.0
TRU-193-0215 105.6 127.9 75.0 - - -
Average 90.0 110.4 62.0 81.3 83.4 54.7
*
Seam strength controls the overall wall strength, except for PIC-138-0091 and SAN-
510-0412.

116
Table 7.5 continued

Design (kN/m) Current (kN/m)


Culvert name
Yield Buckling Seam Yield Buckling Seam
ADA-041-0096 1179 1500 792 965 816 606
ALL-309-1664 1541 1827 1095 1319 1562 909
ALL-696-0410 1319 1665 909 1179 1487 792
ATB-193-2336 1759 2115 1290 1541 1849 1095
ATB-307-1801 1541 1827 1095 1541 1827 1095
CLI-068-0942 1759 2255 1290 1541 1315 1095
COL-170-1981 1541 1922 1095 1319 1090 909
COL-172-1209 1541 1827 1095 1319 1030 909
DEF-018-1560 749 966 420 749 644 420
DEL-037-0333 1319 1103 909 1319 1103 909
FAY-734-0790 1319 1665 909 1179 988 792
FRA-062-2688 1179 1489 792 965 807 606
GAL-218-0578 1759 2126 1290 1541 1862 1095
HIG-134-0782 1319 1665 909 1179 988 792
HOC-093-1975 1319 1624 909 - - -
HOC-664-1172 1319 1624 909 1179 1451 792
JAC-327-0077 1541 1922 1095 1319 1090 909
KNO-013-2162 1319 1624 909 1179 962 792
KNO-062-1480 1541 1957 1095 1319 1114 909
LAK-608-0303 1541 1906 1095 1319 1041 909
LIC-079-2276 749 970 420 749 969 420
LIC-661-0805 1179 1311 792 1179 1311 792
MAD-038-0935 1179 1521 792 965 826 606
MAD-056-1490 1319 1624 909 1179 960 792
MAD-665-1071 1179 1489 792 1179 1487 792
MAR-229-0093 1541 1827 1095 - - -
MAR-309-1248 1179 1521 792 1179 1521 792
MAR-423-0214 1179 1489 792 965 1214 606
MAR-423-0379 1179 1391 792 1179 1391 792
MRW-019-0456 965 1217 606 965 1217 606
MRW-095-1637 1179 1521 792 1179 1521 792
MRW-095-1796 1179 1489 792 1179 1487 792
NOB-083-0419 1319 1544 916 1179 1379 792
PIC-138-0091 749 933 420 447 166* 250
PIC-159-0868 1319 1675 909 1319 1675 909
PIC-752-0748 1179 1375 792 1179 1375 792
SAN-412-0575 1179 1481 792 - - -
SAN-510-0412 1541 1967 1095 1541 1086* 1095
TRU-193-0215 1541 1867 1095 - - -
Average 1313 1611 905 1186 1217 798
*
Seam strength controls the overall wall strength, exceptions: PIC-138-0091 and SAN-
510-0412.

117
7.7.3 Rating factor based on minimum cover
The rating factor based on minimum cover, RFC, is determined from the fill cover
depth, AASHTO minimum cover, and span length (Equation 7.16). RFC is independent
of culvert material properties and thickness of corrugated metal plate as well as internal
or external loads. Once RFC controls the load rating, RF of the designed culvert never
changes unless the cover depth over the culvert is increased, i.e., culvert properties and
damage are irrelevant.
Equation 7.16 is derived based on the plastic moment capacity (Equation 7.18)
suggested by Duncan and Drawsky (1983). Before using Equation 7.16 directly, the
engineer has to check Equation 7.18 using data from the culvert inspection report.
However, the parameters used in Equation 7.18 unfortunately do not affect or are not
related to RFC currently calculated from Equation 7.16. According to the current
equation for RFC (Equation 7.16), the fill cover depth is compared with the AASHTO
specified minimum cover, and the safety factor for the fill cover is compared with the one
for AASHTO minimum cover without considering structural strength and safety. For
example, the culvert SAN-412-0575 has a span of 13.0 ft (3.96 m), a fill depth of 1.5 ft
(457 mm), and a gauge number of 8. RFC of this culvert is 1.1 which allows legal traffic
over the culvert. However, if the culvert is checked using Equation 7.18. H / Sc is 0.1154,

and Fp is 1.55 for operating loads (see Figure 7.8), and k3 is 0.0719 for relative

compaction (RC) of 95 %. Therefore, M p is 5.399 k-ft/ft (24.02 kN-m/m) for RC of

95 %. The required maximum moment strength for gauge number 8 is 4.22 k-ft/ft (18.77
kN-m/m) (Table 7.3). This means that the design gauge number does not satisfy the
required maximum moment. For SAN-412-0575, the design gauge number needs to be 5
( M p _ max = 5.54 k-ft/ft (24.64 kN-m/m)) or smaller. As shown in this example, the design

gauge number does not satisfy the requirement if Equation 7.18 is considered in the RFC
calculation. However, k3 in Equation 7.17 depends on the RC of the soil. According to
ODOT’s Construction and Material Specifications (1951-present), the compaction
requirement is higher than 96%. Therefore, in the example above, k3 may need to be

increased more, and M p may be smaller. Still, Equation 7.16 should not be used without

checking the structural requirements given by Equation 7.18. Although the gauge number

118
of 8 for SAN-412-0575 satisfies every requirement during the initial load rating design, it
does not ensure that RFC of the culvert is always satisfactory because the culvert may
have significant section loss due to steel corrosion and may never develop its maximum
or plastic moment capacity before failure.

7.7.4 Rating factor based on wall strength


Once a culvert is designed and constructed, the thrust forces due to backfill soil,
TE , and live and impact loading, TL +i , in Equation 7.15 can not be changed if the cover

and design load are not changed. Therefore, the wall strength, Tcap , is the only variable

affecting rating factor for wall, RFW. Tcap changes as the culvert condition changes. Thus,

the relationship between RFW and Tcap (Equation 7.15) can be expressed as a linear

function as shown in Figure 7.10. A change in RFW can be checked visually through Tcap .

For example, the design Tcap of ADA-41-0096 is 54.27 k/ft and design RFW is 22.8.

Current RFW and Tcap of the culvert are 41.54 and 15.8, respectively. As shown in Figure

7.11, RFW is reduced by 30% when Tcap drops 24%. If Tcap is reduced by 50%, RFW

will be equal to 8.2, a 63% reduction from the design RFW. Although ODOT’s current
general appraisal number for this culvert is lower than 2, this culvert is safe theoretically.
In order to reduce RFW down to 1.0, the design Tcap has to be decreased by about 85%.

Thus, the current method for RFW seems to be too conservative, at least for this culvert.

119
Figure 7.10: Relationship between RFW and Tcap .

Figure 7.11: RFW example for culvert ADA-41-0096.

7.7.5 Controlling rating factor


Table 7.6 shows RFW, RFC and controlling rating factor, RF for design load
rating as well as current load rating of the 39 culverts tested in this study. All values
reported in the table were calculated and reported by ODOT using CMP-Excel program.
However, the cover depth measured by the researchers in the field was used in the
calculations instead of ODOT’s design cover depth. As shown in the table, the
controlling RF is equal to RFC with the exception of culverts DEL-037-0333 and

120
KNO-062-1480. Both culverts are deep culverts with fill covers more than 16 ft. Usually
RFW for deep culverts is very high because the effect of external live load is small.
According to Equation 7.15, lower values of TL +i correspond to higher RFW. If TL +i is

very low because of deep fill cover, using the available capacity, Tava = Tcap − 1.95TE may

be better than using RF. In this study, there are three deep culverts, CLI-068-0942, DEL-
037-0333, and KNO-062-1480. Their available capacities based on current load ratings
are 41.5 (605.7), 2.3 (10.2), and 23.3 k/ft (103.6 kN/m), respectively.
RFs for some culverts did not change as shown in Table 7.6 although their general
appraisal numbers (GANs) are low or they were built a long time ago. For example,
ATB-307-1801 was built in 1962 and has a GAN of 4. However, its current RF is the
same as its design RF while the culvert has heavy pack of rust and perforations according
to its latest inspection report. Even if its gauge number is changed from 4 to 12, the
overall RF is controlled by RFC and may not change because the fill depth and span
length are still the same.
Table 7.7 shows the overall RF for the worst possible scenario, in which the
gauge number is reduced to 12 (thinnest possible corrugated plate), deflection is 10%
(maximum allowed), and span length increases 2 Rt . The overall RF for deep culverts,
CLI-068-0942, DEL-037-0333, and KNO-062-1480 are negative because their available
capacities are negative as the Tcap is reduced significantly. However, the RFs of other

culverts are positive and they can still support external loads. Although the controlling
RFs of some culverts are lower than 1.0, they are very similar to their design RF. The
design, current, and worst case RF values are plotted in Figure 7.12 ( RF ≤ 18 ). The
results show that the majority of the 39 culverts will always have RFs larger than 1.0 and
will be safe. This suggests that the current load rating procedures need to be revised as
they seem to be unconservative and generally ineffective in identifying possible unsafe
culverts.

121
Table 7.6: Design and current rating factors for 39 test culverts*.

RFW RFC Overall RF


Culvert name
Design Current Design Current Design Current
ADA-041-0096 22.8 15.8 7.3 6.6 7.3 6.6
ALL-309-1664 32.2 25.7 6.5 6.5 6.5 6.5
ALL-696-0410 13.5 11.5 2.8 2.7 2.8 2.7
ATB-193-2336 22.2 18.5 2.8 2.7 2.8 2.7
ATB-307-1801** 24.6 24.6 4.3 4.3 4.3 4.3
CLI-068-0942 469.9 334.1 251.5 227.1 251.5 227.1
COL-170-1981 3.2 2.5 0.4 0.4 0.4 0.4
COL-172-1209 12.1 9.4 1.8 1.6 1.8 1.6
DEF-018-1560 7.5 7.1 4.4 4.0 4.4 4.0
**
DEL-037-0333 15.5 15.5 41.0 41.0 15.5 15.5
FAY-734-0790 7.3 6.0 1.5 1.4 1.5 1.4
FRA-062-2688 30.6 20.8 11.5 10.3 11.5 10.3
GAL-218-0578 38.7 32.2 6.7 6.7 6.7 6.7
HIG-134-0782 10.0 8.1 2.0 1.8 2.0 1.8
HOC-093-1975 23.6 - 5.8 - 5.8 -
HOC-664-1172 11.2 11.0 2.1 2.1 2.1 2.1
JAC-327-0077 14.5 11.2 2.3 2.1 2.3 2.1
KNO-013-2162 47.7 37.7 17.1 15.7 17.1 15.7
KNO-062-1480 359.7 220.7 266.8 241.8 266.8 220.7
LAK-608-0303 9.2 7.2 1.5 1.4 1.5 1.4
LIC-079-2276 10.4 10.3 6.6 6.6 6.6 6.6
LIC-661-0805** 20.4 20.4 6.0 6.0 6.0 6.0
MAD-038-0935 8.1 5.7 2.1 1.9 2.1 1.9
MAD-056-1490 42.8 33.7 14.6 13.1 14.6 13.1
MAD-665-1071 16.9 16.8 4.5 4.4 4.5 4.4
MAR-229-0093 15.8 - 2.2 - 2.2 -
MAR-309-1248 13.1 12.9 3.5 3.4 3.5 3.4
MAR-423-0214 32.9 23.4 12.8 12.5 12.8 12.5
MAR-423-0379 15.2 15.2 3.6 3.6 3.6 3.6
MRW-019-0456 11.2 11.1 3.7 3.6 3.7 3.6
MRW-095-1637** 8.7 8.7 2.3 2.3 2.3 2.3
MRW-095-1796 54.4 53.9 27.5 27.2 27.5 27.2
NOB-083-0419 4.1 3.5 0.7 0.7 0.7 0.7
***
PIC-138-0091 6.8 1.6 3.2 2.6 3.2 1.6
PIC-159-0868 76.1 76.1 35.2 35.2 35.2 35.2
PIC-752-0748** 38.5 38.5 17.1 17.1 17.1 17.1
SAN-412-0575 4.8 - 1.1 - 1.1 -
SAN-510-0412 6.5 6.0 1.1 1.0 1.1 1.0
TRU-193-0215 29.2 - 5.8 - 5.8 -
*
Live load surcharge was not considered in calculations.
**
No change in RF.
***
This value came from reduced section area not from change in gauge number.

122
Table 7.7: Controlling RFs for design, current, and worst case scenario*.

Overall RF
No. Culvert name
Design Current Worst case
1 ADA-041-0096 7.3 6.6 6.0
2 ALL-309-1664 6.5 6.5 5.2
3 ALL-696-0410 2.8 2.7 2.3
4 ATB-193-2336 2.8 2.7 2.3
5 ATB-307-1801 4.3 4.3 3.5
6 CLI-068-0942 251.5 227.1 -51.6
7 COL-170-1981 0.4 0.4 0.4
8 COL-172-1209 1.8 1.6 1.5
9 DEF-018-1560 4.4 4.0 3.5
10 DEL-037-0333 15.5 15.5 -50.9
11 FAY-734-0790 1.5 1.4 1.3
12 FRA-062-2688 11.5 10.3 9.4
13 GAL-218-0578 6.7 6.7 5.4
14 HIG-134-0782 2.0 1.8 1.7
15 HOC-093-1975 5.8 - 4.8
16 HOC-664-1172 2.1 2.1 1.8
17 JAC-327-0077 2.3 2.1 2.0
18 KNO-013-2162 17.1 15.7 9.0
19 KNO-062-1480 266.8 220.7 -103.3
20 LAK-608-0303 1.5 1.4 1.3
21 LIC-079-2276 6.6 6.6 5.5
22 LIC-661-0805 6.0 6.0 2.6
23 MAD-038-0935 2.1 1.9 1.8
24 MAD-056-1490 14.6 13.1 8.7
25 MAD-665-1071 4.5 4.4 3.7
26 MAR-229-0093 2.2 - 1.9
27 MAR-309-1248 3.5 3.4 2.8
28 MAR-423-0214 12.8 12.5 10.1
29 MAR-423-0379 3.6 3.6 3.0
30 MRW-019-0456 3.7 3.6 3.0
31 MRW-095-1637 2.3 2.3 2.0
32 MRW-095-1796 27.5 27.2 13.4
33 NOB-083-0419 0.7 0.7 0.6
**
34 PIC-138-0091 3.2 1.6 1.6**
35 PIC-159-0868 35.2 35.2 15.6
36 PIC-752-0748 17.1 17.1 2.1
37 SAN-412-0575 1.1 - 0.9
38 SAN-510-0412 1.1 1.0 0.9
39 TRU-193-0215 5.8 - 4.6
*
Live load surcharge was not considered in calculations.
**
This value came from reduced section area not from change in gauge number.

123
18
Design RF
15 Current RF
Controlling rating factor
Worst RF
12

3
1.0
0
0 5 10 15 20 25 30 35 40
Culvert number

Figure 7.12: Comparison of controlling rating factors for design, current and worst case
conditions.

124
CHAPTER 8

PROPOSED LOAD RATING PROCEDURE

8.1 Introduction
The purpose of this chapter is to develop an improved load rating procedure and
load rating recommendations based on the field test results, analysis results and
evaluation of current ODOT’s load rating method. As introduced in Chapter 7, rating
factor (RF) of a culvert can be different depending on the engineer who rates the culvert.
In this chapter, recommendations are made to provide consistency for load rating
evaluation and to reduce bias or mistakes. Basic steps involved in the proposed procedure
are shown in Figure 8.1. The development and details of the procedure will be presented
in this chapter.
The load rating procedure proposed in this study is similar to the current ODOT
load rating procedure. Comparison of Figures 7.1 and 8.1 shows that there are a few
differences between the two procedures. First, RF based on minimum cover (RFC) is not
used in the proposed load rating procedure. However, the design cover depth is used to
ensure structural stability. The specified design cover depth must be smaller than the
proposed minimum cover depth, which varies with the design span length and design
gauge number. Second, formulations to calculate the thrust capacity are the same.
However, new reduction factors are introduced to reduce the design seam strength and
wall area to reflect the inspected condition of the culvert. The usage of the buckling
factor is extended for three cases to consider different culvert conditions such as crown
deflection and local buckling. Finally, depending on the magnitude of the thrust produced
by external live loads, two load-rating procedures are proposed. The first one is for
culverts having a thrust due to live load less than 1.0 kips/ft (14.59 kN/m). This is
typically the case in deep culverts because the effect of live load pressure on the crown is
negligible when the backfill height is too large. When the thrust due to live load is very

125
close to zero, RF is infinity because RF is inversely related to live load thrust. Note that
the inventory load rating procedure is not defined for deep culverts because the changes
in live loads do not affect the response of deep culvert. The operating and inventory load
ratings must be calculated when the thrust due to live load is larger than 1.0 k/ft (14.59
kN/m). The RF calculation procedures are the same as those currently specified by
ODOT.

8.2 Basic features of the proposed load rating procedure


In this section, the proposed load rating procedure will be summarized and the difference
between the current ODOT load rating procedure and proposed one will be briefly
discussed. For the most part, the proposed procedure is very similar to the current ODOT
procedure. As discussed in Section 7.7, however, current ODOT load rating method has
some weak points, which need to be addressed. The proposed procedure is developed by
eliminating or modifying parts of the current ODOT procedure identified as deficient in
Chapter 7. In addition, the proposed method recommends two general appraisal numbers
for wall and seam condition to reflect the current culvert condition as reported in the
inspection report (BR-86) so that the engineer who rates the culvert does not need to use
engineering judgment. The basic features of the proposed load rating procedure are
summarized below.

• Rating factor based on minimum cover, RFC is eliminated and replaced with
Section 8.3 of this report.
• Current ODOT RFC and AASHTO minimum cover requirements are independent
of culvert material properties and structural properties. The proposed equation
relates the minimum cover to span length and structural properties.
• Changing the design gauge number based on engineering judgment is not allowed
in the proposed procedure. ODOT’s current RF highly depends on the rating and
gauge number assigned by the engineer during culvert inspection. In the proposed
procedure, once the gauge number is determined during initial design, the design
gauge number will always be same throughout the life of the structure.

126
Was Yes
this culvert rated Check the bridge inspection report (BR-86) if
previously? there is a change in geometry

No
Check the proposed minimum Input basic geometrical and material properties
cover in Table 8.3 for adequacy
of the design gauge number
(See Section 8.3) Check BR-86, and determine the seam strength
reduction rate, φs ( S s' = φs S s ) (Section 8.4.2.1)

Calculate the seam strength, φseam S s'

Determine the area reduction rate, φ A ( A' = φ A A)


(Section 8.4.2.2)

Calculate the wall yield strength, φwall f y A'

Calculate the buckling stress, f b

Any
increase in measured No
buckling or deflection? Calculate the wall buckling strength, φbkl fb A '

Yes
The thrust capacity, Tcap is the smallest of
① Wall yield strength
Calculate a new buckling factor
② Seam strength
(Sections 8.4.3)
③ Wall buckling strength

Yes 1.3TL +i ≤ 1.0 k/ft Calculate the wall thrust due to earth cover, TE
Calculate the wall thrust due to live load, TL +i
?
Determine load rating No
for the culvert. Determine operating load rating (RFo) and
(Section 8.5.1) inventory load rating (RFi) (Section 8.5.2)

Figure 8.1: Proposed load rating flow chart.

127
• New strength reduction factors, which are related to appraisal numbers, are
proposed instead of changing the design gauge number when the culvert is
damaged. Two appraisal numbers are recommended; one for seam strength and
the other for wall area to properly represent the current condition of the culvert.
• A new formula is recommended for buckling factor to consider severe local
buckling because previous buckling factor was related to deflection only, not local
buckling.
• The procedure to calculate the thrust capacity is the same as current ODOT
procedure. However, different formulations are recommended to calculate the
strength reduction factors, which represent current culvert condition.
• For lightly loaded pressure or deep culverts, a different load rating procedure is
recommended. For deep culverts, it is assumed that the live load effect on the
culvert is negligible if the thrust due to live load (1.3TL +i ) is smaller than 1.0
kips/ft (14.6 kN/m) or the cover depth is larger than 6.5 ft (2.0 m).

8.3 Adequacy of the design gauge number for the cover depth
As discussed in Section 7.3.4, current ODOT rating factor based on minimum
cover (RFC) is independent of culvert material properties, thickness of corrugated metal
plate (or gauge number) and internal or external loads. Although the culvert can be
damaged and deteriorated over time, RFC is never changed because RFC is a function of
cover depth over the crown and span length of the culvert. Once the span and cover of the
culvert satisfy the design requirements, ODOT engineers do not need to check RFC again
because RFC is always the same. In addition, according to ODOT Manual of Bridge
Inspection (2006), the purpose of bridge inspection is to manage bridges with accurate
and adequate information. If the updated information from the bridge inspection is not
used for RFC, ODOT engineers do not need to inspect culverts. Thus, the evaluation
method may need to be changed. In this study, current ODOT RFC will not be used.
However, it is very important to ensure that the design cover depth and span length
satisfy the requirements for load rating and structural safety. Unfortunately, there are no
guidelines and code requirements for cover depth, span length, or both. Although
AASHTO Standard (2002) provides a guideline for the minimum cover regarding span

128
length, this is not adequate for the structures damaged due to steel corrosion, section loss,
pitting, or other reasons. For example, if a culvert with the AASHTO minimum cover has
a section loss of 50%, the AASHTO minimum cover is the same as the design cover. The
culvert may need more cover depth to reduce live load pressure on the damaged culvert.
Also, the AASHTO minimum cover is not related to structural properties of culvert. Thus,
in this study a new method is proposed to calculate RFC. The method provides a
relationship between the cover depth, span length, and material properties.

8.3.1 Minimum soil cover


Generally, the external live load effect on the culvert increases as the cover H
decreases. If H is very shallow, the surface live load can cause excessive deflection,
buckling, or severe damage of the culvert. Therefore, determination of minimum depth of
cover is very important. According to AASHTO (2002), the minimum cover level for
culvert with structural plate pipe shall be span length/8 but not less than 12 inches. The
surface live load is transferred to the culvert through pavement and soil. The effect of live
load on the culvert is based on the truncated pyramid model as shown in Figure 8.2
discussed in detail in Section 7.5.3. The procedures for the calculation of live load
pressure on the culvert are based on the assumption that soil is elastic. In Figure 8.2, the
surface load ( w ) represents a tire contact area of width Bs and length Ls . S E is the
spreading factor, and is equal to 1.75 for AASHTO Standard and 1.15 for AASHTO
LRFD where the cover depth ( H ) is 2.0 ft (0.61 m) or greater.
The truncated pressure on the culvert due to the tire contact pressure is
W
q= (8.1)
( Bs + S E H )( Ls + S E H )

where Bs is 10 in. (254 mm) and Ls is 20 in. (508 mm) for the standard AASHTO HS-20

vehicle. If W is known and the critical pressure which causes local buckling or excessive
deflection of the culvert can be evaluated, the minimum cover depth can be found by
solving Equation 8.1. Thus, the minimum cover can be defined as the cover depth of a
safe and stable culvert system subjected to a large number of passes of external live load.

129
W

Figure 8.2: Truncated pyramid showing how a surface load ( W ) spreads over a base area
at cover depth ( H ).

8.3.2 The critical pressure


The critical location in a culvert where the maximum deflection or maximum
moment develops due to external loads is generally near the crown. Tests to failure of
corrugated steel culverts show that static loads over the crown can be many times greater
than the load on one side (Watkins and Anderson, 1999). In order to find the critical
pressure near the crown, some researchers (Timoshenko and Gere, 1961; Meyerhof,
1982; and Moore, 1989) investigated when and where the buckling happens. If a
parabolic arch structure is subjected to a uniform load ( q ) distributed along its span as
shown in Figure 8.3, there will be axial compression but no bending in the arch. If the
intensity of the uniform load is increased, the arch starts to buckle at a critical load ( qcr ),
which can be calculated from Equation 8.2 (Timoshenko and Gere, 1961).
EI
qcr = γ 4 (8.2)
Sc 3

where E is the elastic modulus, I is moment of inertia, and γ 4 is the numerical factor

Hc
depending on the ratio , where H c is the height of the arch. Generally, the top part of
Sc
pipe-arches is similar to a half circle. In this study, it is assumed that the shape of pipe-
arch is a half circle, and each corner of the circle has a fixed or hinged support. γ 4 is 97.4

130
and 38.4 for the fixed and hinged support, respectively.
Once the culvert design is completed, qcr in Equation 8.2 can be calculated

because γ 4 , Sc , E and I are known. If the maximum pressure ( q ) on the crown due to
dead and live load can be determined for a given cover depth, and the calculated pressure
can be compared to qcr to decide whether the design cover depth is adequate. If the

calculated q is larger than qcr , the design cover depth is not sufficient to support the
design load. Then, either the design cover needs to be increased to support the design
load, or the design live load needs to be reduced until q is smaller than qcr .

Figure 8.3: A parabolic arch with hinged and fixed supports subjected to a uniform load.

8.3.3 Proposed minimum cover


The previous sections show that if q is lower than qcr , the design cover depth
satisfies the minimum cover requirement for the load rating evaluation. Thus, the
minimum cover ( h ) to carry the design live load over the culvert can be determined by
setting q equal to qcr . In order to determine h , there can be many variables to be

131
considered. The variables include the span length, culvert height, gauge number or
thickness corrugated steel plate, and dead and live load. In practice, the culvert geometry
is typically selected from the standard layouts. Standard culvert sizes are frequently used
in most designs. The standard detail layouts of pipe-arches investigated in this study are
listed in Table 8.1. In this study, the standard HS-20 design truck is used as the external
live load. The dead load is considered as the soil weight over the culvert crown. The
appropriate load factors for dead and live load are used to ensure the safety and usability
of culvert. The pressure on the crown due to the dead and live load is calculated from the
following equation by modifying Equation 8.1.
W
q = φDγ s H + φL (8.3)
( Bs + 1.75H )( Ls + 1.75H )

where γ s is the soil density, H is the clear cover above the crown, w is the surface load,

and φD and φL are load factors for dead and live load, respectively. When the HS-20

truck is used for the design, Bs is 10 in. (254 mm) and Ls is 20 in. (508 mm). The
pressure on the crown is also a function of cover depth (Equation 8.3).
In this study, the minimum covers are calculated using an iterative procedure, and
the results are listed in Table 8.2. Factors used in the calculations are γ s = 120 lb/ft 3

(18.85 kN/m3), γ 4 = 38.4 , φD = 1.95 and φL = 1.75 . The standard AASHTO HS-20 load

is used to calculate the surface load ( W = 32 kips=142.34 kN ). Table 8.2 shows the
minimum covers for the standard culverts with different gauge numbers or different
corrugated plate thickness. Most of the minimum covers are smaller than the AASHTO
minimum cover requirement. It means that the AASHTO minimum cover is adequate to
use culvert design. However, a few culverts with thin plate thickness and with span
length larger than 14 ft-10 in. (4.52 m) do not satisfy the AASHTO requirement. These
results indicate that the AASHTO requirement may not apply to some large-span culverts
constructed with relatively thin corrugated steel plates.

132
Table 8.1: Pipe-arch size and layout details (18 in. (457 mm) corner radius, Rc)
constructed with 6 × 2 in. (152 × 51 mm) corrugated structural plate (AISI,1994).

Dimensions Waterway Layout dimensions


area,
Span, Sc Height, Hc B Rt Rb
2
ft-in. ft-in. ft in. in. in.
10-3 6-9 55 23.9 5.13 14.86
10-8 6-11 58 26.1 5.41 12.77
10-11 7-1 61 25.1 5.49 15.03
11-5 7-3 64 27.4 5.78 13.16
11-7 7-5 67 26.3 5.85 15.27
11-10 7-7 71 25.2 5.93 18.03
12-4 7-9 74 27.5 6.23 15.54
12-6 7-11 78 26.4 6.29 18.07
12-8 8-1 81 25.2 6.37 21.45
12-10 8-4 85 24.0 6.44 26.23
13-5 8-5 89 26.3 6.73 21.23
13-11 8-7 93 28.9 7.03 18.39
14-1 8-9 97 27.6 7.09 21.18
14-3 8-11 101 26.3 7.16 24.80
14-10 9-1 105 28.9 7.47 21.19
15-4 9-3 109 31.6 7.78 18.90
15-6 9-5 113 30.2 7.83 21.31
15-8 9-7 118 28.8 7.89 24.29
15-10 9-10 122 27.4 7.96 28.18
16-5 9-11 126 30.1 8.27 24.24
16-7 10-1 131 28.7 8.33 27.73

133
Table 8.1 continued

Dimensions Waterway Layout dimensions


area,
Span, Sc Height, Hc B Rt Rb
m m m2 mm mm mm
3.12 2.06 5.11 607.06 130.30 377.44
3.25 2.11 5.39 662.94 137.41 324.36
3.33 2.16 5.67 637.54 139.45 381.76
3.48 2.21 5.95 695.96 146.81 334.26
3.53 2.26 6.22 668.02 148.59 387.86
3.91 2.31 6.60 640.08 150.62 457.96
3.76 2.36 6.87 698.50 158.24 394.72
3.81 2.41 7.25 670.56 159.77 458.98
3.86 2.46 7.53 640.08 161.80 544.83
3.91 2.54 7.90 609.60 163.58 666.24
4.09 2.57 8.27 668.02 170.94 539.24
4.24 2.62 8.64 734.06 178.56 467.11
4.29 2.67 9.01 701.04 180.09 537.97
4.34 2.72 9.38 668.02 181.86 629.92
4.52 2.77 9.75 734.06 189.74 538.23
4.67 2.82 10.13 802.64 197.61 480.06
4.72 2.87 10.50 767.08 198.88 541.27
4.78 2.92 10.96 731.52 200.41 616.97
4.83 3.00 11.33 695.96 202.18 715.77
5.00 3.02 11.71 764.54 210.06 615.70
5.05 3.07 12.17 728.98 211.58 704.34

134
As mentioned earlier in this chapter, the AASHTO minimum cover may not be
applicable to the structures damaged due to steel corrosion. Generally, the plate thickness
of these culverts becomes thinner due to section loss and pitting. Many researchers
(Bellar and Ewing, 1984; Potter et al., 1991; NCHRP 254, 1998) have reported culvert
failure and damage caused by metal loss associated with corrosion and related
degradation. The corroded part of steel plate turns to rust. Rust cannot resist external
loads because it is not strong enough. It is questionable if a culvert with the minimum
cover can support the external load after its plate thickness is reduced significantly. Thus,
the minimum cover requirements should include geometric and structural properties
associated with damage to ensure durability and safety of the culvert during its service
life. In this study, section loss of culvert plate will be considered in minimum cover
calculations.
Potter et al. (1991) reported that in Ohio, most of metal loss due to corrosion
occurs at the waterside in culvert. Although side and bottom plates are usually thicker
than top plates, it is typically assumed that culverts are fabricated with plates with the
same thickness. The thinnest plate used in the culvert governs the design calculations.
The same factors and load, except the moment of inertia, used to calculate the minimum
covers in Table 8.2 are used to determine the proposed minimum cover. In order to
calculate the proposed minimum cover, it is assumed that only the structural property is
changed when the culvert has damage. The structural property used in the calculation is
the moment of inertia of culvert plate. The proposed minimum cover in Table 8.3 is
calculated when the moment of inertia is half of the original value. In the table, the
minimum covers larger than 3 ft are marked as #N/A because it is recommended that the
plate thickness is not adequate for the safety of culvert. If the engineer needs to use the
plate (the gauge number) marked as #N/A in the table, the larger number between the
cover depth larger than 3 ft and the AASHTO minimum cover must to be selected.
The proposed minimum covers in the table show that the AASHTO minimum
cover cannot apply to culverts with number 12 gauge because the minimum cover
requirement proposed in the study is higher than the AASHTO minimum cover
requirement. The thicker plate needs to satisfy the required minimum cover as the span
length of culvert increases. For example, assume that there is a culvert with the design

135
span of 13 ft-11 in. (4.24 m), number 8 gauge, and the cover depth of 2 ft (0.61 m). The
required cover is 2.10 ft (0.64 m) as given in Table 8.3. Therefore, the engineer needs to
increase cover depth or increase the design plate thickness. If the engineer decides to
increase the plate thickness, number 7 gauge may be the best choice because the required
minimum cover is 1.83 ft (0.56 m). If the design span length is not listed in the table, the
engineer can calculated the required minimum cover by interpolating the numbers
between the closest spans in the table.

Table 8.2: The minimum covers in feet for the standard culverts with different gauge
numbers.

Gauge number AASHTO


Span minimum
(ft-in.) cover
1 3 5 7 8 10 12
(ft)
10-3 0.13 0.18 0.24 0.31 0.37 0.49 0.66 1.28
10-8 0.17 0.23 0.29 0.38 0.44 0.56 0.75 1.33
10-11 0.20 0.26 0.33 0.41 0.48 0.61 0.80 1.36
11-5 0.26 0.32 0.40 0.49 0.56 0.70 0.91 1.43
11-7 0.28 0.35 0.42 0.52 0.59 0.74 0.95 1.45
11-10 0.31 0.38 0.46 0.56 0.63 0.79 1.01 1.48
12-4 0.38 0.45 0.53 0.64 0.72 0.89 1.14 1.54
12-6 0.40 0.47 0.56 0.67 0.75 0.92 1.18 1.56
12-8 0.42 0.50 0.58 0.70 0.78 0.96 1.23 1.58
12-10 0.44 0.52 0.61 0.72 0.81 1.00 1.27 1.60
13-5 0.52 0.61 0.70 0.83 0.93 1.13 1.45 1.68
13-11 0.59 0.68 0.79 0.92 1.03 1.26 1.62 1.74
14-1 0.62 0.71 0.82 0.96 1.07 1.30 1.68 1.76
14-3 0.64 0.73 0.85 0.99 1.10 1.35 1.75 1.78
14-10 0.73 0.83 0.95 1.11 1.24 1.51 2.00 1.85
15-4 0.81 0.91 1.05 1.22 1.36 1.68 2.29 1.92
15-6 0.83 0.94 1.08 1.26 1.41 1.74 2.42 1.94
15-8 0.86 0.97 1.11 1.30 1.45 1.80 2.56 1.96
15-10 0.89 1.00 1.15 1.34 1.50 1.86 2.73 1.98
16-5 0.99 1.11 1.27 1.49 1.67 2.13 > 3.0 2.05
16-7 1.01 1.14 1.31 1.53 1.73 2.22 > 3.0 2.07
Bold letters indicate that the numbers are greater than the AASHTO minimum cover.

136
Table 8.2 continued

Gauge number AASHTO


Span minimum
(m) cover
1 3 5 7 8 10 12
(m)
3.12 0.040 0.055 0.073 0.094 0.113 0.149 0.201 0.390
3.25 0.052 0.070 0.088 0.116 0.134 0.171 0.229 0.405
3.33 0.061 0.079 0.101 0.125 0.146 0.186 0.244 0.415
3.48 0.079 0.098 0.122 0.149 0.171 0.213 0.277 0.436
3.53 0.085 0.107 0.128 0.158 0.180 0.226 0.290 0.442
3.91 0.094 0.116 0.140 0.171 0.192 0.241 0.308 0.451
3.76 0.116 0.137 0.162 0.195 0.219 0.271 0.347 0.469
3.81 0.122 0.143 0.171 0.204 0.229 0.280 0.360 0.475
3.86 0.128 0.152 0.177 0.213 0.238 0.293 0.375 0.482
3.91 0.134 0.158 0.186 0.219 0.247 0.305 0.387 0.488
4.09 0.158 0.186 0.213 0.253 0.283 0.344 0.442 0.512
4.24 0.180 0.207 0.241 0.280 0.314 0.384 0.494 0.530
4.29 0.189 0.216 0.250 0.293 0.326 0.396 0.512 0.536
4.34 0.195 0.223 0.259 0.302 0.335 0.411 0.533 0.543
4.52 0.223 0.253 0.290 0.338 0.378 0.460 0.610 0.564
4.67 0.247 0.277 0.320 0.372 0.415 0.512 0.698 0.585
4.72 0.253 0.287 0.329 0.384 0.430 0.530 0.738 0.591
4.78 0.262 0.296 0.338 0.396 0.442 0.549 0.780 0.597
4.83 0.271 0.305 0.351 0.408 0.457 0.567 0.832 0.604
5.00 0.302 0.338 0.387 0.454 0.509 0.649 >0.914 0.625
5.05 0.308 0.347 0.399 0.466 0.527 0.677 >0.914 0.631
Bold letters indicate that the numbers are greater than the AASHTO minimum cover.

137
Table 8.3: The proposed minimum covers in feet for the standard culverts with different
gauge numbers.

Gauge number AASHTO


Span minimum
(ft-in.) cover
1 3 5 7 8 10 12
(ft)
10-3 0.45 0.53 0.62 0.74 0.83 1.02 1.30 1.28
10-8 0.53 0.61 0.71 0.83 0.93 1.14 1.46 1.33
10-11 0.57 0.66 0.76 0.89 1.00 1.22 1.56 1.36
11-5 0.66 0.76 0.87 1.02 1.14 1.38 1.80 1.43
11-7 0.69 0.79 0.91 1.06 1.18 1.44 1.89 1.45
11-10 0.74 0.84 0.96 1.12 1.26 1.54 2.04 1.48
12-4 0.84 0.95 1.09 1.27 1.42 1.75 2.44 1.54
12-6 0.87 0.99 1.13 1.32 1.47 1.83 2.64 1.56
12-8 0.91 1.03 1.17 1.37 1.54 1.92 2.93 1.58
12-10 0.94 1.06 1.22 1.42 1.59 2.01 #N/A 1.60
13-5 1.07 1.21 1.38 1.63 1.84 2.43 #N/A 1.68
13-11 1.19 1.34 1.54 1.83 2.10 #N/A #N/A 1.74
14-1 1.23 1.39 1.60 1.90 2.21 #N/A #N/A 1.76
14-3 1.27 1.44 1.66 1.99 2.33 #N/A #N/A 1.78
14-10 1.43 1.62 1.89 2.34 3.00 #N/A #N/A 1.85
15-4 1.58 1.80 2.14 2.85 #N/A #N/A #N/A 1.92
15-6 1.63 1.87 2.24 #N/A #N/A #N/A #N/A 1.94
15-8 1.69 1.94 2.35 #N/A #N/A #N/A #N/A 1.96
15-10 1.74 2.02 2.48 #N/A #N/A #N/A #N/A 1.98
16-5 1.98 2.35 #N/A #N/A #N/A #N/A #N/A 2.05
16-7 2.05 2.47 #N/A #N/A #N/A #N/A #N/A 2.07
Bold letters indicate that the numbers are greater than the AASHTO minimum cover.

138
Table 8.3 continued

Gauge number AASHTO


Span minimum
(m) cover
1 3 5 7 8 10 12
(m)
3.12 0.137 0.162 0.189 0.226 0.253 0.311 0.396 0.390
3.25 0.162 0.186 0.216 0.253 0.283 0.347 0.445 0.405
3.33 0.174 0.201 0.232 0.271 0.305 0.372 0.475 0.415
3.48 0.201 0.232 0.265 0.311 0.347 0.421 0.549 0.436
3.53 0.210 0.241 0.277 0.323 0.360 0.439 0.576 0.442
3.91 0.226 0.256 0.293 0.341 0.384 0.469 0.622 0.451
3.76 0.256 0.290 0.332 0.387 0.433 0.533 0.744 0.469
3.81 0.265 0.302 0.344 0.402 0.448 0.558 0.805 0.475
3.86 0.277 0.314 0.357 0.418 0.469 0.585 0.893 0.482
3.91 0.287 0.323 0.372 0.433 0.485 0.613 #N/A 0.488
4.09 0.326 0.369 0.421 0.497 0.561 0.741 #N/A 0.512
4.24 0.363 0.408 0.469 0.558 0.640 #N/A #N/A 0.530
4.29 0.375 0.424 0.488 0.579 0.674 #N/A #N/A 0.536
4.34 0.387 0.439 0.506 0.607 0.710 #N/A #N/A 0.543
4.52 0.436 0.494 0.576 0.713 0.914 #N/A #N/A 0.564
4.67 0.482 0.549 0.652 0.869 #N/A #N/A #N/A 0.585
4.72 0.497 0.570 0.683 #N/A #N/A #N/A #N/A 0.591
4.78 0.515 0.591 0.716 #N/A #N/A #N/A #N/A 0.597
4.83 0.530 0.616 0.756 #N/A #N/A #N/A #N/A 0.604
5.00 0.604 0.716 #N/A #N/A #N/A #N/A #N/A 0.625
5.05 0.625 0.753 #N/A #N/A #N/A #N/A #N/A 0.631
Bold letters indicate that the numbers are greater than the AASHTO minimum cover.

8.4 Thrust capacity of wall


As discussed in Sections 7.5.1, 7.6, and 7.7.3, the thrust capacity of wall ( Tcap ) is

determined as the lowest of the factored seam strength ( φseam S s ), wall strength ( φwall f y A ),

and buckling strength ( φbkl fb A ). The proposed procedure to calculate Tcap is very similar

to that currently used by ODOT. However, different reduction factors for seam strength,
wall area, and local buckling are introduced in this study. These factors are related to
general appraisal number reflecting the current condition of the culvert reported in the
BR-86 inspection report. Thus, using the recommended factors can be more effective
than changing the design gauge, or reassigning a new gauge number to account for the

139
observed damage. Engineering judgment is required for appraisal numbers.

8.4.1 General appraisal numbers for seam strength and wall area
When an engineer changes the design gauge number of the culvert (for example,
in ODOT’s CMP-Excel program), the material properties and seam strength of the
corrugation plate are automatically changed. If the culvert has damage in the corrugation
plate and seam connection, changing the gauge number can be reasonable because the
gauge number affects material properties as well as seam strength. However, if the
culvert has corrugation plate damage and section loss due to steel corrosion and pitting,
and does not have seam damage or has minor seam damage not enough to decrease seam
strength, changing gauge number can be ineffective. We propose to decrease section area
and seam strength based on the specified design values without changing the gauge
number. In order to use a reasonable reduction rate for section area and seam strength,
reported appraisal numbers for wall area ( N A ) and seam ( N s ) are used to reflect current
culvert condition in the proposed method. For the 39 culverts tested in this research, the
inspection reports prepared by the authors (OSU research) are compared to ODOT’s
Bridge Inspection Reports (BR-86) as shown in Tables 8.4 and 8.5. Table 8.4 compares
ODOT’s general appraisal number (GAN) to N A as they are related to wall area damage.
Table 8.5 summarizes ODOT’s inspection report and OSU inspection data for the
observed seam damage. Coding method to assign OSU’s GANs follows the ODOT
Manual of Bridge Inspection (2006), which describes ODOT’s coding method and
condition assessment guideline, and is introduced in Appendix D. As shown in Table 8.4,
there are small discrepancies between ODOT’s GANs and OSU numbers for corrugation
wall area. Most of the GAN and N A values for the corrugation section area are very

similar. However, most of the OSU appraisal numbers for seam ( N s ) are higher than
ODOT’s GANs. Thus, we recommend using separate appraisal numbers to report the
wall and seam condition.

140
Table 8.4: Culvert inspection reports for the plate areas of 39 test culverts.

ODOT OSU research


Culvert name
GAN Inspection report NA Inspection report
Looks like this culvert may have been
Top and side areas of rusting from
leakage problems near the top and right
ADA-041-0096 4 leakage with section loss, holes near 4
side. In these areas, some plates are
the top
heavily rusted.
Top of the culvert is repaired with
Rust and efflorescence at the top have
ALL-309-1664 5 6 concrete coating. Bottom plates are
been coated in 2004.
rusted but look good shape.
Bottom sections and the plate
Heavy pack rust on bottom with
ALL-696-0410 5 5 connections at the crown are rusted.
section lost, endwalls spalling.
The crown is defected a little.
Heavy pack rust and loss of section Heavy rust along the top seams and
ATB-193-2336 4 along bottom, perforations. Shape is 4 Extensive steel corrosion in bottom
good. was observed.
Heavy pack rust and many
perforations along both sides under
northerly embankment. No Extensive steel corrosion is observed
overlapping or crushing in these near the entrance. Although these areas
panels. The sections under the are not under roadway, maybe need to
ATB-307-1801 4 5
roadway are in pretty good shape. be coated or strengthened to prevent
There are two sunken areas on the local collapse. The sections under the
north slope above culvert roadway are in good condition.
perforations. These depressions are 1
to 2’ deep.
Some parts of two structural plates at
Bottom with some loss of section. both sides of the culvert are corroded
CLI-068-0942 7 7
Leakage around lower half. lightly. The culvert is in good
condition.
Rust and pitting of isolated areas above
concrete invert. However, overall
Some heavy rust, pitting and much
appearance of the culvert is clean. This
loss of section above new concrete
COL-170-1981 5 5 culvert has a channel flow problem.
invert. Much cracking, spalling out
Scour at inlet is found. There is a gap
and heavy scaling. Scour at inlet rear.
between the concrete wingwall and
culvert inlet.
Heavy rust, pitting and loss of section
mostly in bottom and at invert. Some
Extensive rust and pitting in bottom are
COL-172-1209 5 minor sag. Large open crack at right 5
observed. Top has deflected slightly.
forward and rear through top corners
next to culvert. Some pushing.
Some pack rust on bottom with some
section loss. A few holes rusted Extensive rust and pitting are observed
thought and thin throughout the in bottom. Local bucking exists near
DEF-018-1560 5 5
bottom. Rear side seam on the right outlet. However, the sections under the
side has slight sag with some roadway are in pretty good shape.
separation between panels.

141
Table 8.4 continued

ODOT OSU research


Culvert name
GAN Inspection report NA Inspection report
The shape of the culvert is not
symmetrical because one side has
Moderate deformation throughout top
DEL-037-0333 5 4 settled and deflected more than the
of can, seeping with efflorescence.
other side. Isolated rust on the top is
found.
Bottom plates are rusted. The
Section loss of pipe ends and bottoms connections between plates were
FAY-734-0790 5 5
because of rusting, egg shape. detached and created slight gaps at the
inlet.
Seeping, corrosion rust through at
Rust and pitting at bottom. One side is
FRA-062-2688 5 bottom, slight deformation, egg shape 5
deflected slightly.
at outlet.
Bottom plates have section loss. A localized bulge at the crown is
GAL-218-0578 5 Localized corrosion at top. Perforated 5 found. This area is rusted. Bottom
at midspan. plates are rusted.
The sides along seams have leakage The culvert has a lot of damage
with rusting with section loss. Edges concentrations at the plate connections
HIG-134-0782 5 of plates in several areas are thin with 5 due to corrosion. Worse areas are 6 ft
section rusted away. A few small away from the mid of the culvert.
areas in top have slight cusping (1/8”) These areas are rusted very much.
Heavy rust and pitting are observed in
Localized corrosion at top. Bottom
HOC-093-1975 5 6 bottom. Isolated rust along seams at top
plates heavy deep pitting.
and side.
A lot of cracks and holes as well as
heavy rust and pitting are observed in
bottom. Probably this culvert needs to
HOC-664-1172 3 No record 3
be repaired. There is localized rust
along seam at top but it does not look
bad.
Along the top bolt line approximately
There are extensive deposits of sand
10 ft back from inlet is a 10’ long
and gravel in the culvert. This culvert
rusty area from leakage coming along
JAC-327-0077 5 4 has widespread damage along seams at
the seams with a lot of section loss to
top. The damaged parts are heavily
plates. Top is bowed down and sides
rusted and has leakage problem.
bowed in a little.
The culvert looks very clean. No rust
KNO-013-2162 6 No remarks for this inspection 7
and damage are found.
Inside of the culvert is clean but few
KNO-062-1480 6 No remarks for this inspection 7 isolated slight pitting along seams at
bottom is found.
This culvert has a lot of steel corrosion
Rusting section loss. Some flattening. problems such as rusting, pitting, and
Heavy rusting section loss at the section loss. Most of connections
LAK-608-0303 4 crown seam under roadway. As much 3 between culvert and concrete wall are
as 5 inch of inlet toe wall is exposed heavily rusted. Also, the crown under
due to scouring. roadway has extensive heavy rust, deep
pitting. Water flow is very fast.

142
Table 8.4 continued

ODOT OSU research


Culvert name
GAN Inspection report NA Inspection report
The bottom of the culvert is rusty
LIC-079-2276 6 No remarks for this inspection 7
slightly.
The bottom of the culvert is rusty
LIC-661-0805 6 No remarks for this inspection 7
slightly
Plates at the bottom are corroded
extensively and heavily. Several holes
MAD-038-0935 3 Rust throughout bottom, deformation 3 are found in corner. Near the top, the
seams are detached and created gaps at
several locations.
Isolated corrosion problem at top and
Left side deformation, bottom side under the roadway. Bottom is
MAD-056-1490 6 corroded, Top has rust with section 6 slightly rusted. Overall shape of the
loss. Left rear scour culvert is good. The culvert is slightly
deformed but it looks not bad.
The bottom of the culvert is fairly
heavy rust and moderate pitting. The
Bottom corroded, slight deformation
MAD-665-1071 6 6 culvert is very slightly deformed near
right and let at top.
the mid span of culvert. However,
overall shape is good.
There is heavy rust, deep pitting and
cracks left side and top along seams.
MAR-229-0093 5 No record 5 But, these damages are not under the
roadway. The shape under the roadway
is good. Bottom is slightly rusted.
Slight deformation at center top and The appearance of the culvert seems
MAR-309-1248 7 8
sides. like very clean.
Bottom of the culvert is moderately
Corrosion at bottom with section loss,
MAR-423-0214 6 6 rusted. The culvert is slightly
0.5 inch deformation at top right.
deformed.
The top and side plates along seams are
moderately rusty with section loss.
MAR-423-0379 7 Corrosion at bottom 6
Bottom of the culvert is moderately
rusted.
MRW-019-0456 9 No remarks for this inspection 9 The inside of the culvert is very clean
The inside of the culvert is clean, but
MRW-095-1637 7 Scour at both ends 8
superficial rust on bottom.
The inside of the culvert is clean, but
MRW-095-1796 7 No remarks for this inspection 8
superficial rust on bottom.
The culvert has a lot of damage
concentrations along seams due to steel
corrosion especially near the inlet and
Inlet, outlet, top and sides heavy
NOB-083-0419 4 4 outlet. The concrete walls are damaged
corrosion with section loss.
and broken along vertical cracks above
the crown. Water flows into the crown
through these cracks.

143
Table 8.4 continued

ODOT OSU research


Culvert name
GAN Inspection report NA Inspection report
Localized bulges are found. Heavy rust
and pitting along seams on top. The
culvert is extremely deflected and
PIC-138-0091 2 No record 2
flattened on the left side of the crown.
Openings between plate connections
are found.
The culvert is very clean. However, the
interface between pipe and concrete
PIC-159-0868 8 No remarks for this inspection 8 walls is detached and had visible gaps.
A lot of rust stains are found around
the gaps.
PIC-752-0748 8 No remarks for this inspection 8 The culvert is very clean
The plate connections at top are
Rust on top and sides. Slightly
detached and created gaps mainly
SAN-412-0575 5 depressed. There is slight scour on the 4
because of heavily steel corrosion and
rear side.
pitting.
Rust on the bottom and the lower bolt Bottom sides are detached and created
SAN-510-0412 5 lines. Top depressed about 4 in. and 4 gaps, which allow for leakage of water
0.25 in gaps. from backfill soil.
Rust is forming in many areas. There
Bottom and sides are heavily rusted.
are perforations in the top. There is
There are extensive heavy corrosion
TRU-193-0215 3 water leaking though in these areas 3
damages with section loss and water
that are heavily rusted. Slight
leakage.
deformation at both inlet and outlet.

144
Table 8.5: Culvert inspection reports for the seams of 39 test culverts.

ODOT OSU research


Culvert name
GAN Inspection NS Inspection
Several bolts missing in side and In the top right side, most of blots are
ADA-041-0096 4 bottom, three bolts missing and bolts 5 rusted very much and they can be
very rusty in the top. broken easily.
Top seams and bolts have heavy rust Bolts at the top and right side are rusty
ALL-309-1664 5 and efflorescence and much section 7 but look fine because these are
loss. These have been coated. repaired.
Bolts and nuts in bottom are easily
Bolts and nuts have much section loss
ALL-696-0410 5 5 broken because of heavy rust. Some
on bottom, some area almost rusted.
bolts at the top are rusty.
Seams are pack rusted in some with Most of bolts are rusted heavily and
ATB-193-2336 4 5
minor cusping along top seam. broken easily along the top seams.
Most of seams, bolts and nuts are
rusted from leakage with section loss
Heavy pack rust and many
near the entrance. Although these areas
ATB-307-1801 4 perforations along both sides under 5
are not under roadway, maybe need to
northerly embankment.
be coated or strengthened to prevent
local collapse.
Leakage around lower half bolts
causing rusting and rust stains to Many bolts on both sides are rusted
CLI-068-0942 7 7
multi plates. Left end has isolated area from leakage.
of loss of seams and blots.
Much rust and section loss on some of
Rust and pitting of isolated areas above
the blots above concrete invert. Much
COL-170-1981 5 6 concrete invert. However, overall
cracking, spalling out and heavy
appearance of the culvert is clean.
scaling.
Most of bolts and nuts in bottom are
Heavy rust, pitting and loss of section
COL-172-1209 5 6 rusted. Some nuts at top are marginally
mostly in bottom and at invert.
rusted.
Some pack rust on bottom with some Heavily rust and pitting are observed in
DEF-018-1560 5 5
section loss. bottom.
Bolts on top near the inlet are rusted.
DEL-037-0333 5 No remarks for the seam and blots 6
Overall shapes are good.
Bolts on sides are moderately rusted.
Section loss of pipe ends and bottoms Some bolts near inlet and sides are
FAY-734-0790 5 6
because of rusting, egg shape. slightly rusted. There are seam
openings near inlet.
Rust and pitting at bottom. Some bolts
FRA-062-2688 5 Seam rust through at bottom. 6 at sides are slightly rusted. There are
slight openings on seam at corners.
There are seam openings on top. Bolts
Bottom plates have section loss.
GAL-218-0578 5 5 at sides are heavily rusted. Bolts on
Localized corrosion at top.
bottom are easily broken.

145
Table 8.5 continued

ODOT OSU research


Culvert name
GAN Inspection NS Inspection
The sides along seams have leakage
with rusting with section loss. Seams
There are seam openings on top. Blots
running down one nut missing in
and nuts on the rusted area near mid
forward wall left of center. Forward
HIG-134-0782 5 5 span are heavily corroded with section
wall has several unusual bolts that
loss. Rust is easily broken by touching
were used when pipe was installed.
it with hands.
Rusting is a little worse in left half of
pipe.
Heavy rust and pitting are observed in
Localized corrosion at top. Bottom
HOC-093-1975 5 6 bottom. Bolts on bottom are heavily
plates heavy deep pitting.
rusted.
A lot of cracks and holes as well as
HOC-664-1172 3 No record 3 heavy rust and pitting along seams are
observed in bottom.
A 10’ long rusty area from leakage
along the seams with a lot of section This culvert has widespread damage
loss to plates. One nut missing at right along seams at top. Bolts and nuts on
JAC-327-0077 5 5
forward and one bolt is loosing in the the damaged parts are heavily rusted
side at right forward. Some water with section loss and leakage problem.
dripping down around blots.
The culvert looks very clean. Bolts and
KNO-013-2162 6 No remarks for this inspection 8
nuts are very tight and no openings.
The culvert looks very clean. Bolts and
KNO-062-1480 6 No remarks for this inspection 8
nuts are very tight and no openings.
This culvert has a lot of steel corrosion
problems such as rusting, pitting, and
Eight 3/4 inch to 1 inch diameter
section loss. Most of bolts and nuts
rusting through holes. Heavy rusting
LAK-608-0303 4 4 near the connections between culvert
section loss with water infiltration at
and concrete wall are heavily rusted.
the crown seam under roadway.
Heavy seam damages on top under
roadway are observed.
The culvert looks very clean. Bolts and
LIC-079-2276 6 No remarks for this inspection 8
nuts are very tight and no openings.
The culvert looks clean. Some nuts are
LIC-661-0805 6 No remarks for this inspection 7
slightly rusted.
Seam openings near top and at both
sides are observed. Blots and nuts at
the bottom are corroded extensively
MAD-038-0935 3 Rust throughout bottom, deformation 3
and heavily. In these areas, several
cracks and holes are also found in
bottom corner.
Left side deformation, bottom There are seam openings at both sides.
MAD-056-1490 6 corroded, Top has rust with section 6 Bolts and nuts on the isolated rust areas
loss. Left rear scour on top are heavily rusted.
There are minor joint openings on top
Bottom corroded, slight deformation
MAD-665-1071 6 7 and sides, but it looks good. Bolts on
right and let at top.
bottom are heavily rusted.

146
Table 8.5 continued

ODOT OSU research


Culvert name
GAN Inspection NS Inspection
Seam openings on top under the
roadway are observed. There are cracks
MAR-229-0093 5 No record 4
along the blot line on the left side and
top near inlet.
The appearance of the culvert seems
Slight deformation at center top and
MAR-309-1248 7 7 like very clean. There are minor seam
sides.
openings.
Corrosion at bottom with section loss, Bolts on bottom are moderately rusted.
MAR-423-0214 6 7
0.5 inch deformation at top right. Joints and seams are loose slightly.
Bolts on top are moderately rusty with
section loss. Bolts on bottom are also
MAR-423-0379 7 Corrosion at bottom 6
moderately rusted. There are slight
seam openings on top and sides.
MRW-019-0456 9 No remarks for this inspection 9 The inside of the culvert is very clean
The inside of the culvert is clean, but
MRW-095-1637 7 Scour at both ends 7
minor joint openings at both sides.
Bolts on bottom are moderately rusted.
MRW-095-1796 7 No remarks for this inspection 8 Bolts on top and sides are slightly
rusted.
Most of bolts are heavily rusted along
seams near inlet and outlet due to water
Inlet, outlet, top and sides heavy leakage on top. However, seams, bolts
NOB-083-0419 4 4
corrosion with section loss. and nuts under roadway are not bad,
but there is moisture along joints
because of leakage.
Bolts and nuts along seams on top are
heavily rusted. This culvert is deflected
PIC-138-0091 2 No record 3
because of large seam openings on
sides and top.
PIC-159-0868 8 No remarks for this inspection 8 The culvert is very clean.
PIC-752-0748 8 No remarks for this inspection 8 Seams and joints are tight
The plate connections at top are
Rust on top and sides. Slightly detached and created gaps mainly
SAN-412-0575 5 depressed. There is slight scour on the 4 because of heavily steel corrosion and
rear side. pitting. Most of bolts and nuts are
heavily rusted on these areas.
Seam openings at both sides and
Rust on the bottom and the lower bolt
corners are found. There are heavy
SAN-510-0412 5 lines. Top depressed about 4 in. and 4
rusting and pitting with leakage at both
0.25 in gaps.
corners.
Many are pack rusted and there are Bottom and sides are heavily rusted.
TRU-193-0215 3 perforations. Some of the blot heads 3 Bolts and nuts are heavily rusted in
are rusted off. these areas. It looks very serious.

147
8.4.2 Reduction rate for seam strength and wall area
8.4.2.1 Seam strength reduction rate
The tension or compression capacity at longitudinal seams between adjacent
corrugated plate sections can be determined by calculating the shear resistance of the
bolted connections. Shear force applied on the longitudinal bolted joints larger than the
shear resistance can cause bolt failure, or cracking distortion of seam plate connections.
Plate cracked from bolt to bolt on seams may not support the design load, and cause
excessive deflections or failure of the culvert.
In this study, a new reduction factor ( φS ) is introduced to account for damage and

deterioration that can occur in seams of existing culverts. The new reduction factor φS is

related to the appraisal number ( N S ) for seam strength, which was discussed in the

previous section. To determine the seam strength reduction rate φS as a function of

appraisal number for seam strength N S , the relation between N S and φS needs to be
established. Degler et al. (1988) investigated about 900 corrugated metal pipe-arch
structures in Ohio to identify and determine the cause of numerous problems observed
with these structures. According to their research, cracking along the seams has no
correlation with the age of the structure, and significant distortion or sagging of the
structure influences seam cracking. Only a very few culverts (about 0.2 %) among 900
culverts experienced major cracking (corresponding to N S = 3 ), seam condition of most
culverts (about 99.6 %) was better than the generally fair condition (corresponding to
N S ≥ 5 ) regardless the age of culvert. Usually, culverts were removed, replaced or

repaired before N S was reduced to less than 2. Thus, in this study, it is assumed that the

poor condition ( N S = 3 ) corresponding to 50% reduction in seam strength ( φS = 0.50 ).

The proposed φS values shown in Table 8.6 are based on this assumption.

The modified seam strength ( S s' ) can then be expressed as

S s' = φS S s (8.4)

The seam strength S s was discussed in Section 7.4.1. Thus, thrust due to the
modified seam strength is

148
Ts = φseam S s' = φseamφs S s (8.5)
Table 8.6: Proposed seam strength reduction rate for corrugated metal pipe arches.

Inspected φS
NS Shape Seams and joints
condition (%)
9 New condition 100
Good, smooth curvature;
8 Good condition span within 3% greater Tight, no openings 100
than design
Minor flattening of
Minor joint or seam openings
Generally good bottom;
7 Potential for backfill 95
condition Span within 3-5% greater
infiltration
than design
Fair, smooth curvature in Minor cracking at bolts
top half, bottom flat; Evidence of backfill
6 Fair condition 90
Span within 5% greater infiltration through seams or
than design joints
Moderate cracking at bolt
Fair, significant distortion
holes along one seam near
Generally fair in top in one location;
5 bottom 80
condition bottom has slightly
Deflection of pipe due to
reverse curvature
backfill infiltration
Marginal, significant
Moderate cracking at bolt
distortion along top;
Marginal holes on one seam near top
4 bottom has reverse curve; 65
condition Deflection caused by loss of
span more than 7% greater
backfill thought open joints
than design
Poor, extreme deflection
in top arch; bottom has
3 inch long cracks at bolt
3 Poor condition reverse curvature; span 50
holes on one seam
more than 7% greater than
design
Critical, extreme
Critical deflection along top of Plate cracked from bolt to bolt
2 30
condition pipe; span more than 7% on one seam
greater than design
1 Closed to traffic Partially collapsed Failed 0
0 Closed to traffic Collapsed Failed 0

149
8.4.2.2 Area reduction rate
In this research, a new factor ( φ A ) is introduced to account for the reduction in
cross-sectional area as the metal thickness decreases due to corrosion or other effects.
The proposed area reduction rate φ A is related to the general appraisal number for wall

area ( N A ). It was very difficult to find a relationship between φ A and N A because we


could not find research specifically investigating damage, area reduction and strength. In
this study, φ A is considered to be a function of average steel thickness loss typically

associated with corrosion. φ A is defined as the ratio of reduced thickness to initial


thickness.
According to NCHRP 254 (1998), average rates of metal loss govern the design.
In addition, it is assumed that metal loss of culverts due to corrosion occurs not on the
soil side but waterside because Potter et al. (1991) reported that, in Ohio, most of the
metal loss is waterside corrosion. Bellair and Ewing (1984) reported metal loss rate
measured from core samples taken along the inverts of uncoated metal culverts in New
York. According to their experimental research results, the average metal loss rate of
uncoated corrugated steel culverts is about two mils (thousandths of an inch) (0.051 mm)
per year although this rate varies depending on environmental conditions. For example, if
there is a culvert constructed with number 12 gauge (with an uncoated thickness of
0.1046 in. (2.657 mm)), the total metal loss of the culvert will occurs after about 52 years.
However, at the present time, uncoated plates are rarely used any more. Thus, the service
life of coated plates with number 12 gauge (with 0.111 in. (2.819 mm) coated thickness)
should be longer than 52 years. Generally, the desired minimum service life of
galvanized corrugated metal culverts is about 50 years (Potter et al., 1991). It is not easy
to know how much time is needed to destroy the coating of plate. Researchers (Stavors,
1984; Townsend and Zoccola, 1979) experimented on corrosion losses of galvanized
sheets exposed to the atmosphere, and showed that corrosion loss can reach to the coating
film (about 0.5 mils (0.127 mm)) after about 10 years. Therefore, in this study, it is
assumed that the coated film of corrugated plates will be removed approximately 10
years after construction, and then two mils of the plate thickness will be reduced every
year. For example, a culvert with number 12 gauge (0.111 in. (2.819 mm) thick) designed

150
for the service life of 50 years will start to corrode after 10 years when the coating film
will is damaged (remaining thickness is 0.1046 in. (2.657 mm)). At the end of the
remaining service life (40 years), φ A of the culvert can be calculated as 0.24.

(0.1046 in. − 2 mils/year × 40 years)


φA = = 0.24 = 24%
0.1046 in.
This means that only 24% of the plate is available to support the external loads at the end
of the service life. If number 1 gauge, instead of number 12 gauge, is used in the above
example, φ A will increases to 0.71. Thus, φ A of number 1 gauge is 47% larger than that
of the gauge number 12. In other words, the thicker the plate more durable it is.
In this study, the gauge numbers are classified into two parts to provide a range
for φ A . The first category includes thinner plates (from gauge number 12 to 7), and the

other is for thicker plates (from gauge number 5 to 1). As shown in Table 8.7, φ A for the

thinner plate is reduced from 100% to 20%, while φ A for the thicker plates varies
between 100% and 60% depending on the general appraisal number or the condition of
culvert.
The modified wall area can then be expressed as
A' = φ A A (8.6)
Thus, the thrust capacity provided by the wall strength is
Tw = φwall f y A' = φwallφ A f y A (8.7)

where φwall is the capacity modification factor for wall, and is equal to 1.0, f y is the

specified minimum yield strength of the wall plate (See Section 7.4.2).

151
Table 8.7: Area reduction rate for corrugated metal pipe arches.

NA Inspected φA (%)
Shape Steel
condition 1-5* 7-12*
9 New condition 100 100
Good, smooth curvature;
Superficial rust, no
8 Good condition span within 3% greater 100 100
rusting
than design
Minor flattening of
Generally good bottom; Moderate rust
7 95 90
condition Span within 3-5% greater Slight pitting
than design
Fair, smooth curvature in
top half, bottom flat; Fairly heavy rust
6 Fair condition 90 80
Span within 5% greater Moderate pitting
than design
Fair, significant distortion
Generally fair in top in one location; Scattered heavy rust
5 85 65
condition bottom has slightly reverse Deep pitting
curvature
Marginal, significant Extensive heavy rust
distortion along top; Deep pitting
Marginal
4 bottom has reverse curve; Heavy loss of section 80 50
condition
span more than 7% greater Easily punch a hole
than design thru metal
Poor, extreme deflection in
top arch; bottom has Extensive heavy rust
3 Poor condition reverse curvature; span Deep pitting 75 35
more than 7% greater than Scattered perforations
design
Critical, extreme deflection
Critical along top of pipe; span Extensive perforations
2 60 20
condition more than 7% greater than due to rust
design
1 Closed to traffic Partially collapsed
0 Closed to traffic Collapsed
* These numbers indicate the gauge number.

152
8.4.3 Reduction factors for deflection and local buckling
In addition to Equation 7.1, a new formula was proposed to calculate the buckling
factor. The proposed formula is used to consider the effect of local buckling because the
Equation 7.1 can be used only for crown deflection. As discussed in Sections 7.3 and
7.4.3, buckling factor ( φbkl ) is used as a function of the deflection ratio ( δ ) to reduce the
buckling strength of culvert when the culvert has vertical deflection or flattening at the
crown ( Δ ) (see Equation 7.1). However, culvert deflections do not always occur
vertically at the crown. Local buckling and associated deflections do not necessarily
occur at the crown. In this study, local buckling is defined as the ratio of change in the
m
top radius ( = , see Figure 7.4) because buckling is directly associated with the change
Rt

in the geometry. The top radius, Rt is the radius from the center of top curve to the

neutral surface of the cross section of the wall. The suggested φbkl in this study is given in
the following equation
⎧1.0 for no deflection or buckling
⎪0.95 − 5.6δ for deflection at the crown
φbkl = ⎪⎨ (8.8)
⎪0.95 − 4.6 m
for local buckling
⎪⎩ Rt
As explained in 7.4.3, the maximum allowable deflection ratio is 10% for load rating.
The buckling factor is 0.39 when δ equals 0.1. The maximum allowable local buckling
ratio is 12%, and the lowest buckling factor is 0.40.
The equation to calculate the buckling stress ( f b ) is the same as Equations 7.5 and
7.6 (Section 7.4.3). However, the wall area in the buckling strength calculation is
changed to Equation 8.6. Thus, the reduced thrust capacity against buckling is
Tb = φbkl fb A' = φbklφ A fb A (8.9)

8.4.4 Comparison the proposed thrusts to current ODOT thrusts


The thrust capacity of wall ( Tcap ) is the smallest of the three different physical

strengths ( Ts , Tw , and Tb ) as discussed above. The equation for Tcap can be expressed as

153
⎧Ts =φseam S s' = φseamφs S s

Tcap = minimum of ⎨Tw =φwall f y A' = φwallφ A f y A (8.10)

⎩Tb =φbkl f b A = φbklφ A fb A
'

Table 8.8 shows thrusts calculated using current ODOT procedure and the proposed
procedure for the 39 culverts tested in this research. The results obtained from ODOT
procedure are very similar to those determined from the proposed procedure. However,
thrusts calculated from the proposed procedure are somewhat smaller.

8.5 Proposed load rating


Current ODOT rating factor (RF) procedures have two different RFs; one for the
wall strength (RFW) and the other for soil cover (RFC) as explained in Section 7.6.1. In
this study, RFC is not used to evaluate RF of culvert as described in Section 8.2. Instead
of RFC, a minimum cover is calculated and used only during the initial design stage.
Only RFW will be discussed in this section.

8.5.1 Live load effect on the crown


In this study, we suggest a new load rating method for culverts subjected to very
small live load stresses because the current RF of these culverts is very large. Generally,
the effect of external live load over the culvert is reduced as the cover depth increases. If
the denominator (1.3TL +i ) in Equation 7.15 is very close to zero, RFW will be infinity. As
given in Table 7.6, design RFW of culverts CLI-068-0942 and KNO-062-1480 are larger
than 350 because of very small live load stress applied on these deep culverts. A RFW of
350 means that the culvert can support 350 HS-20 vehicles theoretically (total weight is
about 11,200 kips). The RFW can be larger, if the denominator is reduced more. In order
to correct this problem for deeper culverts or culverts subjected to very small live load
pressures, the effect of TL +i is assumed to be negligible in this study. In other words, the
numerator in Equation 7.15 is considered as RFW when the denominator is less than 1.0
kip/ft or the cover depth is larger than 6.5 ft (2.0 m). RFW can then be expressed as
RFW = Tcap − 1.95TE (8.11)

154
Table 8.8: Comparison of the proposed thrusts to current ODOT thrusts.

Current ODOT thrusts (k/ft) Proposed thrusts (k/ft)


Culvert name
Seam Wall Buckling Seam Wall Buckling
ADA-041-0096 41.5 66.1 55.9 43.4 40.4 34.2
ALL-309-1664 62.3 90.4 107.0 71.3 95.0 112.5
ALL-696-0410 54.3 80.8 101.9 49.8 58.8 74.1
ATB-193-2336 75.0 105.6 126.7 70.8 102.4 122.9
ATB-307-1801 75.0 105.6 125.2 60.0 89.7 105.8
CLI-068-0942 75.0 105.6 90.1 84.0 114.4 97.4
COL-170-1981 62.3 90.4 74.7 67.5 89.7 73.9
COL-172-1209 62.3 90.4 70.6 67.5 89.7 69.6
DEF-018-1560 28.8 51.3 44.1 23.0 33.4 28.6
DEL-037-0333 62.3 90.4 75.6 56.1 58.8 59.9
FAY-734-0790 54.3 80.8 67.7 56.1 58.8 49.2
FRA-062-2688 41.5 66.1 55.3 48.8 52.5 44.0
GAL-218-0578 75.0 105.6 127.6 70.8 102.4 123.8
HIG-134-0782 54.3 80.8 67.7 49.8 58.8 49.2
HOC-093-1975* - - - 56.1 72.3 89.1
HOC-664-1172 54.3 80.8 99.4 31.2 31.6 38.9
JAC-327-0077 62.3 90.4 74.7 60.0 89.7 74.2
KNO-013-2162 54.3 80.8 65.9 62.3 81.3 66.4
KNO-062-1480 62.3 90.4 76.3 75.0 100.3 84.7
LAK-608-0303 62.3 90.4 71.3 52.5 84.5 69.2
LIC-079-2276 28.8 51.3 66.4 28.8 46.2 59.8
LIC-661-0805 54.3 80.8 89.8 51.6 72.7 80.8
MAD-038-0935 41.5 66.1 56.6 32.6 40.4 34.6
MAD-056-1490 54.3 80.8 65.8 56.1 72.3 59.0
MAD-665-1071 54.3 80.8 101.9 51.6 64.7 81.5
MAR-229-0093* - - - 60.0 89.7 106.4
MAR-309-1248 54.3 80.8 104.0 51.6 80.8 104.0
MAR-423-0214 41.5 66.1 83.2 51.6 64.7 81.4
MAR-423-0379 54.3 80.8 95.3 48.8 64.7 76.2
MRW-019-0456 41.5 66.1 83.3 41.5 66.1 83.3
MRW-095-1637 54.3 80.8 104.2 51.6 80.8 104.2
MRW-095-1796 54.3 80.8 101.9 54.3 80.8 101.9
NOB-083-0419 54.3 80.8 94.5 43.6 58.8 68.8
PIC-138-0091 17.1 30.6 11.4 14.4 10.3 4.9
PIC-159-0868 62.3 90.4 114.8 62.3 90.4 114.8
PIC-752-0748 54.3 80.8 94.2 54.3 80.8 94.2
SAN-412-0575* - - - 38.0 52.5 66.0
SAN-510-0412 75.0 105.6 74.4 52.5 89.7 75.9
TRU-193-0215* - - - 45.0 84.5 102.4
*
CMP-Excel file is removed from ODOT’s inventory. There is no clue to calculate or
assume the thrust values.

155
Table 8.8 continued

Current ODOT thrusts (kN/m) Proposed thrusts (kN/m)


Culvert name
Seam Wall Buckling Seam Wall Buckling
ADA-041-0096 605.6 964.7 815.8 633.4 589.6 498.6
ALL-309-1664 909.2 1319.3 1561.5 1040.5 1386.4 1641.8
ALL-696-0410 792.4 1179.2 1487.1 726.8 858.1 1081.4
ATB-193-2336 1094.5 1541.1 1849.0 1033.2 1494.4 1793.6
ATB-307-1801 1094.5 1541.1 1827.2 875.6 1309.1 1544.0
CLI-068-0942 1094.5 1541.1 1314.9 1225.9 1669.5 1421.4
COL-170-1981 909.2 1319.3 1090.2 985.1 1309.1 1078.5
COL-172-1209 909.2 1319.3 1030.3 985.1 1309.1 1015.7
DEF-018-1560 420.3 748.7 643.6 335.7 487.4 417.4
DEL-037-0333 909.2 1319.3 1103.3 818.7 858.1 874.2
FAY-734-0790 792.4 1179.2 988.0 818.7 858.1 718.0
FRA-062-2688 605.6 964.7 807.0 712.2 766.2 642.1
GAL-218-0578 1094.5 1541.1 1862.2 1033.2 1494.4 1806.7
HIG-134-0782 792.4 1179.2 988.0 726.8 858.1 718.0
HOC-093-1975* - - - 818.7 1055.1 1300.3
HOC-664-1172 792.4 1179.2 1450.6 455.3 461.2 567.7
JAC-327-0077 909.2 1319.3 1090.2 875.6 1309.1 1082.9
KNO-013-2162 792.4 1179.2 961.7 909.2 1186.5 969.0
KNO-062-1480 909.2 1319.3 1113.5 1094.5 1463.8 1236.1
LAK-608-0303 909.2 1319.3 1040.5 766.2 1233.2 1009.9
LIC-079-2276 420.3 748.7 969.0 420.3 674.2 872.7
LIC-661-0805 792.4 1179.2 1310.5 753.0 1061.0 1179.2
MAD-038-0935 605.6 964.7 826.0 475.8 589.6 504.9
MAD-056-1490 792.4 1179.2 960.3 818.7 1055.1 861.0
MAD-665-1071 792.4 1179.2 1487.1 753.0 944.2 1189.4
MAR-229-0093* - - - 875.6 1309.1 1552.8
MAR-309-1248 792.4 1179.2 1517.8 753.0 1179.2 1517.8
MAR-423-0214 605.6 964.7 1214.2 753.0 944.2 1187.9
MAR-423-0379 792.4 1179.2 1390.8 712.2 944.2 1112.1
MRW-019-0456 605.6 964.7 1215.7 605.6 964.7 1215.7
MRW-095-1637 792.4 1179.2 1520.7 753.0 1179.2 1520.7
MRW-095-1796 792.4 1179.2 1487.1 792.4 1179.2 1487.1
NOB-083-0419 792.4 1179.2 1379.1 636.3 858.1 1004.1
PIC-138-0091 249.6 446.6 166.4 210.2 150.3 71.5
PIC-159-0868 909.2 1319.3 1675.4 909.2 1319.3 1675.4
PIC-752-0748 792.4 1179.2 1374.7 792.4 1179.2 1374.7
SAN-412-0575* - - - 554.6 766.2 963.2
SAN-510-0412 1094.5 1541.1 1085.8 766.2 1309.1 1107.7
TRU-193-0215* - - - 656.7 1233.2 1494.4
*
CMP-Excel file is removed from ODOT’s inventory. There is no clue to calculate or
assume the thrust values.

156
Figure 8.4 shows thrust variation as a function of cover depth. The curved lines in the
figure are thrusts due to live load, and the straight lines are thrusts due to dead load. As
shown in the figure, thrust due to live and dead load increases as span length ( Sc )
increases as a given cover depth. Obviously, as cover increases, the effect of thrust due to
live load decreases.

Cover depth (m)


0 1 2 3 4 5 6
5
70
Sc = 10 ft
4 Sc = 12 ft 60
Sc = 15 ft
Sc = 20 ft 50
Thrust (kip/ft)

Thrust (kN/m)
3 Sc = 30 ft
40
Dead load
Live load
2 30

20
1
10

0 0
0 5 10 15 20
Cover depth (ft)

Figure 8.4: Relationship between thrust due to dead or live load versus cover depth.

8.5.2 Operating and inventory load rating procedures


The procedure to calculate rating factors for operating and inventory load rating
are the same as current ODOT procedure if 1.3TL +i is larger than 1.0 k/ft (14.59 kN/m) as

explained in Section 8.4.1. The rating factor for operating loads ( RFo ) is calculated as

Tcap − 1.95TE
RFo = if 1.3TL +i > 1.0 k/ft (14.59 kN/m) (8.12)
1.3TL +i

157
The rating factor for inventory ( RFi ) is determined as

Tcap − 1.95TE
RFi = if 2.17TL +i > 1.0 k/ft (14.59 kN/m) (8.13)
2.17TL +i
In the proposed procedure, operating and inventory RF are identical for very deep
culverts or culverts under very low live loads, TL +i .

The proposed load rating procedures can be summarized as:


• First, check to make sure the design span, cover and gauge number of the culvert
are adequate for the proposed minimum cover.
• Second, check the inspection report (BR-86) of the culvert, and assign the
appraisal numbers for wall and seam. Find the reduction rates, φs and φ A
corresponding to the assigned appraisal number. Then, calculate the seam thrust,
Ts and wall yield strength, TA .

• Third, calculate the buckling stress, fb and determine the buckling factor based

on BR-86. Then, calculate the wall buckling thrust, Tb .

• Determine the thrust capacity ( Tcap ), which is the smallest of the wall yield thrust ,

seam thrust, and wall buckling thrust calculated above.


• Calculate the wall thrusts due to cover and live load.
• Check whether the live load thrust is small enough to ignore the load rating
calculation (1.3TL +i < 1.0 k/ft (14.59 kN/m)). If the live load effect is very small,
use Equation 8.10 and the operating and inventory RF are identical. Otherwise,
use Equation 8.12 for operating RF and Equation 8.13 for inventory RF.

8.6 Application of proposed load rating


In this section, the proposed load rating procedure will be illustrated using an
example. One of the culverts tested in this research, HOC-664-1172 is selected as an
example. The design span length and height of the culvert are 14 ft-3 in. × 8 ft-11 in.
(4.34 m × 2.72 m), respectively. This culvert is fabricated with number 7 gauge plate and
its cover depth is 2.82 ft (0.86 m). The design gauge number and span length satisfy the
proposed minimum cover as well as the AASHTO minimum cover (Table 8.3). The

158
culver of this culvert does not need to be checked again as the minimum cover
requirement is met. According to the current inspection report, N A = 3 (from Table 8.4)

and N S = 3 (from Table 8.5). Therefore, φs and φ A are 50% and 35% from Tables 8.6

and 8.7, respectively. There is no deflection or local buckling ( φbkl = 1.0 ) reported. Based

on these factors, physical strengths of the culvert are calculated as ( f b calculation is


omitted)
⎧Ts =φseam S s' = 0.67 × 0.5 × 93 = 31.2 k/ft=455.3 kN/m

Tcap = minimum of ⎨Tw =φwall f y A' = 1.0 × 33 × 0.35 × 2.739 = 31.6 k/ft=461.2 kN/m

⎩Tb =φbkl f b A = 1.0 × 40.6 × 0.35 × 2.739 = 38.9 k/ft=567.7 kN/m
'

Thus, thrust due to seam strength ( Tcap = Ts ) controls the load rating of the culvert. TE

and TL +i of the culvert are 2.42 (35.32) and 3.47 k/ft (50.64 kN/m) from equations 7.9
and 7.13, respectively. Equation 8.10 will be used to calculate RF of the culvert because
TL +i is larger than 1.0 k/ft (14.59 kN/m). The operating RF of the culvert is

Tcap − 1.95TE 31.2 − 1.95 × 2.42


RFo = = = 5.9
1.3TL +i 1.3 × 3.47
Based on the result, this culvert can support the design load. However, the appraisal
numbers for wall and seam are very low. Thus, this culvert needs to be monitored
carefully to prevent failure.
The culvert, PIC-138-0091 is used as another example. The design span length
and height of the culvert are 13 ft-5 in. (4.09 m) and 8 ft-5 in. (2.57 m), respectively.
Number 12 gauge and 3.0 ft (0.91 m) cover are used in design. According to Table 8.3,
the design gauge number should not be used for this culvert. Thus, for this culvert, cover
depth should be increased, gauge number should be increased (probably to number 10
gauge), or the allowable live load on the culver needs to be reduced.

8.7 Comparison current ODOT’s load rating to the proposed load rating
In this study, current RFs of the 39 test culverts are calculated based on the
proposed load rating procedures. The proposed RFs are compared to current ODOT RFs.
Table 8.9 shows RFs calculated from ODOT’s load rating procedure and the proposed

159
load rating procedure for the 39 culverts tested in this study. In general, RFs determined
from the proposed method are close to ODOT’s RFs. However, for the deep culverts like
CLI-068-0942 ( H = 22.25 ft (6.78 m), DEL-037-0333 ( H = 16.0 ft (4.88 m)), and KNO-
062-1480 ( H = 24.5 ft (7.47 m)), ODOT’s RFs are much larger than the proposed RFs
because the effect of external live load is very small.
The overall RF for the worst possible scenario, in which the reduction factors are
the minimum possible values, is given for each test culvert in Table 8.10. Most of the
proposed RFo values are smaller than ODOT’s RFW values for the worst case. RFs of
some culverts calculated from the proposed procedure are negative, while ODOT’s RFs
are negative only for deep culverts. This suggests that the proposed load rating
procedures seem to be less conservative and more effective to identify current condition
of culvert. Most of the proposed RFo for the worst case are slightly less than 1.0 as
shown in Figure 8.5. However, some culverts with less than number 5 gauge plates are
strong enough to have RFo higher than 1.0 even for the worst case. The plate thicknesses
of these culverts are much thicker than needed. One of the objectives of the proposed
procedure was to identify culverts if they become unsafe (RF < 1.0). The data presented
in Table 8.10 and Figure 8.5 show that no matter how bad the condition of the culvert is,
most of the culverts tested in this study will always have an RF larger than 1.0, i.e., they
will always be safe. It is possible that these culverts can be unsafe (while RF > 1.0) if the
total vertical crown deflection ratio is larger than 0.10 or the local bucking ratio is larger
than 0.12 (Section 8.4.3).

8.8 Load rating procedures based on AASHTO LRFD Specifications


The design method for bridges has recently been changed from AASHTO
Standard Specifications to AASHTO LRFD Bridge Design Specifications (2006).
Although the load rating procedure and resistance factors are not changed, the method for
live load application has changed. In this section, AASHTO LRFD (2006) method is
briefly discussed and RF based on AASHTO LRFD is compared to the proposed RF.

160
Table 8.9: Comparison of the proposed RF to current ODOT RF.

Current ODOT RFW*** Current OSU RFo***


Culvert name Design RFW
Seam Wall Buckling Seam Wall Buckling
ADA-041-0096 22.8* 15.8 26.7 22.2 16.7 15.3 12.6
*
ALL-309-1664 32.2 25.7 39.6 47.8 30.1 41.8 50.5
#
ALL-696-0410 13.5* (17.6 ) 11.5 17.6 22.4 10.5 12.5 16.1
* #
ATB-193-2336 22.2 (19.5 ) 18.5 26.7 32.3 17.3 25.8 31.3
ATB-307-1801 24.6 24.6 35.8 43.0 19.0 30.0 33.9
CLI-068-0942 468.5* 344.1 564.4 447.6 53.3 83.7 66.6
#
COL-170-1981 3.2 (12.9 ) 2.5 3.6 3.0 2.7 3.6 2.9
#
COL-172-1209 12.1 (15.8 ) 9.4 14.0 10.7 10.2 13.9 9.9
DEF-018-1560 7.5 7.1 13.4 11.4 5.4 8.4 7.0
DEL-037-0333 15.5 15.5 64.2 38.6 13.0 15.7 16.9
#
FAY-734-0790 7.3 (11.4 ) 6.0 9.1 7.6 6.2 6.5 5.4
FRA-062-2688 30.6 20.8 36.1 29.4 25.3 27.6 22.3
#
GAL-218-0578 38.7 (16.4 ) 32.2 47.1 57.8 30.1 45.5 53.0
#
HIG-134-0782 10.0 (13.2 ) 8.1 12.4 10.3 7.4 8.8 7.3
HOC-093-1975 23.6 - - - 10.0 13.1 16.3
#
HOC-664-1172 11.2 (12.8 ) 11.0 16.9 21.0 5.9 6.0 7.6
#
JAC-327-0077 14.5 (14.5 ) 11.2 16.7 13.6 10.8 16.5 13.5
KNO-013-2162 47.7** 37.7 62.0 48.4 45.1 62.4 48.8
KNO-062-1480 359.7* 220.7 457.2 338.9 38.9 64.2 48.6
#
LAK-608-0303 9.2 (9.2 ) 7.2 10.6 8.2 6.0 9.8 8.0
LIC-079-2276 10.4 10.3 19.7 26.0 10.4 17.6 23.3
LIC-661-0805 20.4 20.4 33.7 38.2 19.0 29.7 33.7
#
MAD-038-0935 8.1 (8.1 ) 5.7 9.3 7.9 4.4 5.5 4.7
MAD-056-1490 42.8 33.7 54.9 42.9 35.1 48.1 37.4
MAD-665-1071 16.9 16.8 25.8 32.9 15.9 23.0 26.0
MAR-229-0093 15.8 - - - 12.4 19.1 22.9
#
MAR-309-1248 13.1 (4.8 ) 12.9 19.7 25.6 12.3 19.7 25.6
MAR-423-0214 32.9* 23.4 40.7 52.8 30.5 39.7 51.5
MAR-423-0379 15.2* 15.2 23.9 28.7 13.4 18.6 22.4
MRW-019-0456 11.2* 11.1 18.5 23.7 11.1 18.5 23.7
#
MRW-095-1637 8.7 (14.6 ) 8.7 13.2 17.1 8.2 13.2 17.1
MRW-095-1796 54.4 53.9 87.8 114.7 42.2 68.7 89.8
* #
NOB-083-0419 4.1 (16.3 ) 3.5 5.4 6.3 2.8 3.8 4.5
*
PIC-138-0091 6.8 3.0 6.5 1.6 2.3 1.3 -0.1
PIC-159-0868 76.1* 76.1 119.0 156.3 49.8 77.9 102.3
PIC-752-0748 38.5 38.5 67.0 81.3 35.9 62.5 75.9
**
SAN-412-0575 2.7 - - - 1.9 2.6 3.3
#
SAN-510-0412 6.5 (6.5 ) 6.0 8.5 6.0 4.2 7.2 6.1
TRU-193-0215 29.2** - - - 16.1 33.4 41.2
*
This value is obtained from the ODOT CMP-Excel sheet.
**
This value is based on parameters measured by the authors.
***
Live load surcharge is not considered for the calculation of load rating.
#
The specified design minimum cover is used, and for all others measured cover is used.

161
Table 8.10: Controlling RF for the worst case *.

No. Culvert name ODOT’s worst RFW OSU’s worst RFo


1 ADA-041-0096 8.0 0.8
2 ALL-309-1664 5.2 5.0
3 ALL-696-0410 4.3 0.9
4 ATB-193-2336 4.0 4.8
5 ATB-307-1801 4.5 4.4
6 CLI-068-0942 -51.6 -5.7
7 COL-170-1981 0.9 0.8
8 COL-172-1209 2.7 2.7
9 DEF-018-1560 5.7 0.3
10 DEL-037-0333 -50.9 -26.2
11 FAY-734-0790 2.4 0.6
12 FRA-062-2688 9.6 -0.5
13 GAL-218-0578 5.8 7.2
14 HIG-134-0782 3.3 0.8
15 HOC-093-1975 6.2 0.7
16 HOC-664-1172 3.4 0.7
17 JAC-327-0077 3.7 3.4
18 KNO-013-2162 9.0 -4.9
19 KNO-062-1480 -103.3 -15.3
20 LAK-608-0303 2.4 2.2
21 LIC-079-2276 8.0 0.3
22 LIC-661-0805 2.6 -3.9
23 MAD-038-0935 3.2 0.7
24 MAD-056-1490 8.7 -3.1
25 MAD-665-1071 6.1 0.8
26 MAR-229-0093 3.4 3.3
27 MAR-309-1248 5.1 1.0
28 MAR-423-0214 10.1 -0.9
29 MAR-423-0379 4.2 -0.4
30 MRW-019-0456 5.4 0.4
31 MRW-095-1637 3.4 0.8
32 MRW-095-1796 13.4 -5.5
33 NOB-083-0419 1.2 0.3
34 PIC-138-0091 1.6 -0.1
35 PIC-159-0868 15.6 -4.9
36 PIC-752-0748 2.1 -12.8
37 SAN-412-0575 1.9 0.3
38 SAN-510-0412 1.9 1.6
39 TRU-193-0215 5.6 6.0
*
Live load surcharge was not considered in calculations.

162
16

14 ODOT's RFW for the worst case

Rating factor based on the wall


Proposed RFo for the worst case
12

10

0
0 5 10 15 20 25 30 35 40
Culvert number

Figure 8.5: Comparison of ODOT’s RFW to proposed RF for the worst possible culvert
condition.

8.8.1 Design parameters


The total factored force effect is given by
Q = ∑ηiγ i Qi (8.14)

where ηi is load modifier, γ i is load factor, and Qi is force effects from loads. Load

factors for live load are γ L = 1.75 for culvert design, and γ L = 1.0 for load rating. The

load factor for dead load ( γ D ) varies from 0.90 to 1.95, but γ D = 1.95 is used for culvert
design and load rating (AASHTO LRFD Section 3.4.1).
The maximum live load effects are determined by considering each possible
combination of number of loaded lanes multiplied by the multiple presence factors ( mm )
given in Table 8.11. The factors in the table are based on average one-way daily truck
traffic (ADTT) of 5000 vehicles. A reduction factor ( RV ) may be applied to the multiple

presence factors depending on the average daily volume of traffic. Values for RV are
based on ADTT as follows (AASHTO LRFD Section 3.6.1.1.2):

163
⎧ RV = 0.95, for 100 ≤ ADTT ≤ 1000
⎨ (8.15)
⎩ RV = 0.90, for ADTT < 100
The dimensions of the distributed area at depth is equal to the dimensions of the
tire footprint area at the surface plus the cover depth ( H ) times the soil spreading factor
( S E ). In the AASHTO LRFD Specification, the distribution of surface live load pressure
at depth is spread through the soil cover as follows (AASHTO LRFD Section
C3.6.1.2.6):
⎧ S E = 1.0, for average backfill
⎨ (8.16)
⎩ S E = 1.15, for select granual backfill
The force effects of dynamic vehicle loading on structures is reflected by applying
a dynamic allowance factor ( IM ) to the design load. For buried substructures (e.g.,
culverts), the effects of dynamic loading dissipate with depth as follows (AASHTO
LRFD Section 3.6.2.2):
IM = 33(1.0 − 0.125 H ) ≥ 0% (8.17)
where H is cover depth over the culvert crown in feet units. The force effects of vertical
earth pressure on a culvert are calculated based on a prism load. The prism load is the
weight of the rectangular prism of soil above the culvert, or the free field geostatic stress
at the crown of the culvert multiplied by the culvert span length. If a culvert deflects
vertically, the load supported by the culvert is reduced, i.e., positive arching occurs.
Conversely, negative arching and an increase in the load supported by a culvert occurs
when the side fill settles less than the culvert can deflect vertically. The condition of
equal settlement between the culvert and adjacent side fill is called neutral arching. The
approximate boundaries of soil backfill mobilized by positive, neutral and negative
arching are illustrated in Figure 8.6.

Table 8.11: Multiple presence factor (AASHTO LRFD Table 3.6.1.1.2-1).

Number of loaded lanes Multiple presence factor


1 1.20
2 1.00
3 0.85
>3 0.65

164
Positive arching Neutral arching Negative arching
(VAF < 1) (VAF = 1) (VAF > 1)

Figure 8.6: Soil block mobilized by arching over culverts (FHWA HI-98-032, 2001).

The prism load is represented by the neutral arching. The soil-structure interaction factor
( FE ) is usually considered as the vertical arching factor ( VAF ) (AASHTO LRFD
Section 12.10.2.1).
Resistance factors for seam, wall area, and buckling are the same as AASHTO
Specification. Wall area ( φwall ) and buckling factor ( φbkl ) are 1.0, and seam factor ( φseam )
is 0.67 (AASHTO LRFD Section 12.5.5).

8.8.2 Pressure on the crown due to live load


The live load model in the AASHTO LRFD Specification (2006) is based on
evaluation of various vehicle configurations. The group of vehicles and loadings are
analyzed to determine their force effect on structures. The combination of the design
tandem with uniform lane load and the combination of the HS20 design truck with the
design lane load are useful to determine critical loads for bridge design. The design
tandem consists of a pair of 25.0 kip (111.2 kN) axles spaced at 4.0 ft (1.22 m) apart with
a transverse spacing of 6.0 ft (1.83 m). The design lane load consists of a load, uniformly

165
distributed in the longitudinal direction equal to 0.64 k/ft (9.34 kN/m) with a transverse
spacing of 10 ft (3.05 m). The extreme effect of this load combination is designated as
HL-93 (AASHTO LRFD Section 3.6.1.2.1).
For the design vehicle, the surface tire contact area equals the width of each wheel
( ww = 20 in. (508 mm)) times the wheel length ( wl = 10 in. (254 mm)). Pressure on the
tire contact area is distributed on the culvert through the cover depth. The distributed area
calculation becomes more difficult when distributed areas begin to overlap from adjacent
wheel loads, or adjacent axle loads. For the design truck, the distributed length, LD , due
to the single lane contribution is calculated as:
LD = wl + S E H (8.18)

where LD is length dimension in direction of travel for distributed area at cover depth H ,

and S E is the spreading factor. For the design tandem (in ft units),

4.0 − wl
LD = 2( wl + S E H ) if H ≤ (8.19)
SE
4.0 − wl
LD = wl + S E H + 4.0 if H > (8.20)
SE

The distributed width, WD , due to the single lane contribution for both the design
truck and design tandem is:
WD = 2( ww + S E H ) (8.21)

The distributed area, AD , is then determined by:

AD = LDWD (8.22)
The general relationship for distributing design vehicle live loads to culverts is:
⎛ IM ⎞
mm RV γ L AL ⎜1 + ⎟
DTP = ⎝ 100 ⎠
(8.23)
AD

where DTP is distributed truck or tandem pressure at cover depth H , mm is multiple

presence factor, RV is reduction factor based on ADTT (ADTT = number of trucks per

day in one direction averaged over design life of structure (vehicles/day)), γ L is live load
factor for the limit state evaluated, AL is 32 kips (142.3 kN) for the design truck

166
(AASHTO LRFD 3.6.1.2.2) and 25 kips (111.2 kN) for design tandem (AASHTO LRFD
3.6.1.2.3), and AD is sum of individual distributed wheel pressure areas from total axle
group or net area defined by perimeter of overlapping distributed wheel areas.
The factored pressure distribution due to the uniform land load is:
0.64mm RV γ L
DLP = (8.24)
10.0
The dead load due to vertical earth load is:
γ DWE = γ D FE γ S H (8.25)

where WE is unit load on culvert due to vertical earth pressure, β D is the dead load factor,

FE is soil-structure interaction factor or arching factor, γ S is unit weight of soil above

culvert, SC is diameter or span of culvert.


The factors used in LRFD RF calculations are listed in Table 8.12 and Table 8.13
presents current TL +i and RF calculated from the proposed procedure and AASHTO
LRFD procedure for the 39 test culverts. As shown in the table, most LRFD RFs are
higher than proposed RFs because the effect of external load for LRFD design is higher.
Figure 8.7 shows the overall RF for the worst possible scenario. Most LRFD RFs are less
than proposed RFs. Once LRFD RFs are close to 1.0, the difference between LRFD RFs
and proposed RFs becomes small. Thus, the evaluation of load rating using AASHTO
LRFD Specification is also recommended.

167
8

7 LRFD RF

Rating factor for the worst case


Proposed RF
6

0
0 5 10 15 20 25 30 35 40
Culvert number

Figure 8.7: Comparison of LRFD RFs to proposed RFs for the worst possible culvert
condition.

168
Table 8.12: The factors used in LRFD RF calculations for the 39 test culverts.

H Number
Culvert name mm ADTT RV SE VAF IM
(ft) (m) of Lanes
ADA-041-0096 4.0 1.219 1 1.2 1900 1.00 1.15 1 0.165
ALL-309-1664 5.3 1.615 1 1.2 12420 1.00 1.15 1 0.111
ALL-696-0410 2.6 0.792 1 1.2 1900 1.00 1.15 1 0.223
ATB-193-2336 3.3 1.006 1 1.2 3904 1.00 1.15 1 0.194
ATB-307-1801 4.3 1.311 1 1.2 2150 1.00 1.15 1 0.153
CLI-068-0942 22.3 6.797 1 1.2 5338 1.00 1.15 1 0.000
COL-170-1981 0.9 0.274 1 1.2 3307 1.00 1.15 1 0.293
COL-172-1209 2.6 0.792 1 1.2 1970 1.00 1.15 1 0.223
DEF-018-1560 2.8 0.853 1 1.2 2930 1.00 1.15 1 0.215
DEL-037-0333 16.0 4.877 1 1.2 1780 1.00 1.15 1 0.000
FAY-734-0790 1.8 0.549 1 1.2 570 0.95 1.15 1 0.256
FRA-062-2688 5.3 1.615 1 1.2 9550 1.00 1.15 1 0.111
GAL-218-0578 5.0 1.524 1 1.2 570 0.95 1.15 1 0.124
HIG-134-0782 2.1 0.640 1 1.2 1312 1.00 1.15 1 0.243
HOC-093-1975 4.3 1.311 1 1.2 1900 1.00 1.15 1 0.153
HOC-664-1172 2.5 0.762 1 1.2 1600 1.00 1.15 1 0.227
JAC-327-0077 2.5 0.762 1 1.2 1880 1.00 1.15 1 0.227
KNO-013-2162 7.4 2.256 1 1.2 5070 1.00 1.15 1 0.025
KNO-062-1480 24.5 7.468 1 1.2 3650 1.00 1.15 1 0.000
LAK-608-0303 2.0 0.610 1 1.2 2670 1.00 1.15 1 0.248
LIC-079-2276 3.3 1.006 1 1.2 860 0.95 1.15 1 0.194
LIC-661-0805 5.7 1.737 1 1.2 2390 1.00 1.15 1 0.095
MAD-038-0935 1.9 0.579 1 1.2 3224 1.00 1.15 1 0.252
MAD-056-1490 6.8 2.073 1 1.2 1830 1.00 1.15 1 0.050
MAD-665-1071 3.3 1.006 1 1.2 1520 1.00 1.15 1 0.194
MAR-229-0093 3.0 0.914 1 1.2 1690 1.00 1.15 1 0.206
MAR-309-1248 2.5 0.762 1 1.2 4050 1.00 1.15 1 0.227
MAR-423-0214 5.5 1.676 1 1.2 4160 1.00 1.15 1 0.103
MAR-423-0379 4.0 1.219 1 1.2 4160 1.00 1.15 1 0.165
MRW-019-0456 3.0 0.914 1 1.2 1351 1.00 1.15 1 0.206
MRW-095-1637 2.0 0.610 1 1.2 6600 1.00 1.15 1 0.248
MRW-095-1796 8.2 2.499 1 1.2 4570 1.00 1.15 1 0.000
NOB-083-0419 1.5 0.457 1 1.2 320 0.95 1.15 1 0.268
PIC-138-0091 3.0 0.914 1 1.2 1480 1.00 1.15 1 0.206
PIC-159-0868 8.9 2.713 1 1.2 2870 1.00 1.15 1 0.000
PIC-752-0748 9.0 2.743 1 1.2 1172 1.00 1.15 1 0.000
SAN-412-0575 1.5 0.457 1 1.2 1040 1.00 1.15 1 0.268
SAN-510-0412 1.4 0.427 1 1.2 602 0.95 1.15 1 0.272
TRU-193-0215 4.6 1.402 2 1.0 8530 1.00 1.15 1 0.140

169
Table 8.13: Thrust due to live load and current rating factors for the 39 test culverts.

1.3 TL +i LRFD TL +i
Culvert name Proposed RF LRFD RF
(k/ft) (kN/m) (k/ft) (kN/m)
ADA-041-0096 2136 31173 3488 50904 14.3 9.1
ALL-309-1664 2028 29596 3282 47897 30.1 18.3
ALL-696-0410 4017 58624 6875 100333 10.5 6.6
ATB-193-2336 3723 54333 6109 89154 17.3 10.5
ATB-307-1801 2719 39681 4439 64782 19.0 11.4
CLI-068-0942 129 1883 207 3021 53.3 51.8
COL-170-1981 18383 268280 26510 386884 2.7 2.5
COL-172-1209 5282 77085 9040 131929 9.9 6.9
DEF-018-1560 3031 44234 5127 74823 5.4 3.7
DEL-037-0333 425 6202 676 9865 13.0 11.8
FAY-734-0790 6776 98888 10309 150449 5.4 4.5
FRA-062-2688 1518 22154 2457 35857 22.3 14.4
GAL-218-0578 2076 30297 3203 46744 30.1 19.0
HIG-134-0782 5485 80048 9099 132790 7.3 5.0
HOC-093-1975 2424 35376 3958 57763 10.0 13.0
HOC-664-1172 4842 70664 8337 121669 5.9 3.9
JAC-327-0077 4477 65337 7709 112504 10.8 7.2
KNO-013-2162 1047 15280 1619 23628 45.1 29.7
KNO-062-1480 113 1649 183 2671 38.9 37.2
LAK-608-0303 6596 96261 10808 157731 6.0 4.5
LIC-079-2276 2412 35200 3760 54873 10.4 6.5
LIC-661-0805 2182 31844 3508 51195 19.0 10.4
MAD-038-0935 5361 78238 8666 126471 4.4 3.5
MAD-056-1490 1227 17907 1928 28137 35.1 22.6
MAD-665-1071 2927 42716 4803 70095 15.9 9.6
MAR-229-0093 4394 64126 7342 107148 12.4 7.3
MAR-309-1248 3549 51794 6110 89169 12.3 7.9
MAR-423-0214 1438 20986 2319 33843 30.5 18.3
MAR-423-0379 3042 44395 4966 72473 13.4 8.1
MRW-019-0456 3388 49444 5662 82631 11.1 6.5
MRW-095-1637 5110 74575 8372 122180 8.2 5.8
MRW-095-1796 778 11354 1185 17294 42.2 34.6
NOB-083-0419 11912 173843 17702 258341 2.8 2.3
PIC-138-0091 3547 51765 5927 86498 -0.1 -0.1
PIC-159-0868 655 9559 1004 14652 49.8 48.4
PIC-752-0748 964 14069 1480 21599 35.9 23.1
SAN-412-0575 9109 132936 14249 207949 1.9 2.6
SAN-510-0412 8849 129141 13021 190027 4.2 3.9
TRU-193-0215 2116 30881 2874 41943 16.1 12.7

170
CHAPTER 9

SUMMARY AND CONCLUSIONS

The main objective of this study was to evaluate and improve ODOT’s current
load rating procedures for corrugated metal culverts. This objective is achieved by
carrying out experimental tests of 39 in-service culverts under static and dynamic loads,
by evaluating the response of test culverts using theoretical methods and numerical
simulations, and by reviewing, evaluating and improving the current design practices and
load rating methods. The experimental program was conducted to investigate the
influence of several parameters on the field performance of culverts in Ohio. These
parameters included backfill height, various static and dynamic load applications, and
existing condition, size, shape, and other properties of culverts. Some of the test culverts
were analyzed using a two-dimensional finite element program. Experimental data and
available theoretical methods were used to evaluate the load rating procedures adapted by
ODOT. Recommendations were developed to improve the current load rating procedures.
The conclusions of this research are summarized below.

z Maximum deflections measured in test culverts decreased nonlinearly with increasing


backfill height. Maximum deflections were nearly zero for deep culverts with backfill
heights greater than 13 ft (4 m). Under both static and dynamic loading, culvert
deflections and strains were increased significantly when the backfill height was less
than about 6.5 ft (2.0 m). This suggests that the live load effect on the culvert is
negligible when the cover depth is larger then 6.5 ft (2.0 m) although current
AASHTO Standard neglects the live load effect when the cover depth is larger then 8
ft (2.4 m).

171
z Maximum deflections measured during dynamic truck loading were approximately 10
to 30% less than corresponding maximum deflections for static truck loading.
Experimental data suggest that developing and calibrating culvert design
methodologies using static deflection data is more conservative, and it eliminates the
need to perform more sensitive dynamic load tests.

z There was no significant direct relationship between the maximum deflections and
total truck weight for the culverts tested. However, a more reasonable relationship was
obtained between the maximum deflections and AASHTO equivalent line load, q,
which is a function of backfill height. Additional scatter is observed in the data for q ≥
1.7 kips/ft (25 kN/m). This indicates that culvert behavior is more difficult to predict
when backfill heights are shallower because other factors, such as soil type and culvert
age and condition, likely play a significant role. Culvert deflections decreased with
improved culvert condition (i.e., by increased general appraisal number).

z Deflections predicted by the computer program CANDE for both the silty sand (SM)
and silty clayey sand (SC) and theoretical methods were larger than experimental
results. However, experimental thrust forces are similar to predicted and theoretical
thrust forces. Thrust forces for two different soils (SM and SC) are similar. This result
suggests that the effect of soil type on thrust forces is negligible. As supported by this
conclusion, generally, soil type is not considered in load rating methods.

z ODOT’s current load rating procedures have some deficiencies. First, engineering
judgment can influence the load rating significantly. Two similar culverts may have
the same general appraisal number and similar damage, but may have quite different
load ratings depending on the engineer who rates the culverts. Second, the rating
factor based on minimum cover (RFC) is only a function of backfill depth, AASHTO
minimum cover and span length. Culvert material properties, magnitude of applied
loads or severe damage in culvert do not affect RFC. Finally, for deep culverts or
lightly loaded culverts, the rating factor based on wall (RFW) is significantly large. If

172
the thrust due to live load plus impact is close to zero (no live load pressure on the
crown) RFW will be equal to infinity.

z Rating factors (RF) are calculated for the 39 test culverts assuming that all culverts are
severely damaged and they are in worst possible condition in which the reduction
factor for deflection and the plate thickness are the minimum values. Most of the
calculated RF values were larger than 1.0 indicating that most culverts would be safe
even if they are severely damaged. This suggests that the current load rating
procedures need to be revised as they seem to be unconservative and generally
ineffective in identifying possible unsafe culverts.

z Several recommendations are made to improve the current ODOT load rating
procedures. First, the proposed procedure does not include rating factor for cover
depth, RFC. However, the cover depth is required to be checked during the initial
design stage to ensure structural stability. Formulations to calculate the thrust capacity
are the same. However, new reduction factors are introduced to reduce the design
seam strength and wall area to reflect the current condition of the culvert reported
during annual inspections. The usage of the buckling factor is extended to three cases
to consider different culvert conditions including excessive crown deflection and local
buckling. Finally, two load-rating procedures are proposed depending on the
magnitude of the thrust produced by external live loads. The first one is for culverts
subjected to very small live load stress, or for deep culverts, which have the cover
depth larger than 6.5 ft (2.0 m) or the thrust due to live load is less than 1.0 k/ft (14.6
kN/m). It is recommended that the effect of live load be neglected in load rating
evaluation of deep culverts, and inventory load rating is not necessary. When the
thrust due to live load is larger than 1.0 k/ft (14.6 kN/m), the operating and inventory
rating factors need to be calculated.

z Our analysis showed that the AASHTO minimum cover requirements might not apply
to some large-span culverts constructed with thinner corrugated steel plates. In
addition, the AASHTO minimum cover may not be applicable to damaged structures,

173
for example damaged due to steel corrosion, because the requirement is not related to
structural properties. The proposed minimum cover requirement is related to cover
depth, span length, and structural properties.

z In this study, it is recommended to use separate appraisal numbers for wall and seam
conditions because the condition of wall and seam is not always the same. New
reduction factors are introduced to account for damage and deterioration of seam and
wall of the culvert using the appraisal numbers reported during annual inspections.

z The thrust capacity of wall calculated using the current ODOT procedure is very
similar to the one proposed in this study. However, the wall thrust capacity calculated
from the proposed procedure is typically smaller because the wall, seam and buckling
strengths are reduced more by the proposed new reduction factors.

z We recommend a new load rating method for culverts subjected to very small live
load stresses. Generally, these are deep culverts because the effect of external live load
over the culvert is reduced as the cover depth increases. The rating factors are very
large for these culverts if the current ODOT procedure is used. In the proposed method,
the effect of live load stress is ignored if the cover depth is larger than 6.5 ft (2.0 m) or
the thrust due to live load is less than 1.0 k/ft (14.59 kN/m).

z For the worst possible culvert condition (i.e., the reduction factors have the minimum
possible values for each culvert), proposed rating factors (RF) are smaller than
ODOT’s RFs and are also less than 1.0 for most culverts. This suggests that the
research-proposed load rating procedure is less conservative and more effective in
evaluation of the existing condition of culverts.

z Some culverts with very thick plates (number 5 or smaller gauge number) are strong
enough to have a RF larger than 1.0 even for the worst culvert condition. One of the
objectives of the proposed procedure was to identify culverts if they become unsafe
(when RF < 1.0). No matter how bad the condition of the culvert is, some of the

174
culverts tested in this study will always have a RF larger than 1.0. IN OTHER
WORDS, these culverts will always be safe unless excessive deflections (deflection
ratio at crown ≥ 10%), extreme local buckling or other specified severe damage
conditions are reported.

z The design method for bridges has recently been changed from AASHTO Standard
Specifications to AASHTO LRFD Bridge Design Specifications (2006). Although the
load rating procedure and resistance factors are not changed, the method for live load
application has changed. Based on the results of our load rating analyses using the
AASHTO LRFD Specifications, we recommend to use the AASHTO LRFD
Specifications for live load application, which is more complicated than the method
included in the AASHTO Standard Specifications.

175
LIST OF REFERENCES
AASHTO. 1992. Standard Specifications for Highway Bridges, American Association for
Transportation and Highway and Transportation Officials, Washington, DC.

AASHTO. 1994. Manual for condition evaluation of bridges, 2nd Edition., American
Association for Transportation and Highway and Transportation Officials,
Washington, D.C.

AASHTO. 2000. AASHTO LRFD Bridge Design Specifications. American Association


of State Highway and Transportation Officials, Washington, D.C.

AASHTO. 2002. Standard Specifications for Highway Bridges, 17th Edition. American
Association for Transportation and Highway and Transportation Officials,
Washington, DC.

AISI. 1994. Handbook of Steel Drainage & Highway Construction Products. Fifth
Edition, American Iron and Steel Institute. Washington D. C.

Alan, F. R. 1990. Experimental and Numerical Investigation of a Deep-Corrugated Steel,


Box-Type Culvert. M.S. Thesis. Ohio University, Athens.

Abdel-Sayed, G. 1978. Stability of Flexible Conduits Embedded in Soil. Canadian


Journal of Civil Engineering, Vol. 5, No.3, pp. 463-468.

Abdel-Sayed, G., and Bakht, B. 1981. Soil-Steel Structure Design by the Ontario Code:
Part 2, Structural Considerations. Canadian Journal of Civil Engineering, Vol.
8, No. 3, pp. 545-550.

Adedapo, A. A. 2007. Pavement Deterioration and PE Pipe Behaviour Resulting from


Open-cut and HDD Pipeline Installation Techniques. Ph.D dissertation,
University of Waterloo, Canada.

Akgül, F., and Frangopol, D. M. 2004. Bridge Rating and Reliability Correlation:
Comprehensive Study for Different Bridge Types. Journal of Structural
Engineering, Vol.130, No.7, pp.1063-1074.

Bacher, A. E., and Kirkland, D. E. 1985. California Department of Transportation


Structural Steel Plate Pipe Culvert Research: Design Summary and
Implementation. Transportation Research Record 1008, TRB, Washington D.
C., pp. 89-94.

Bacher, A. E., and Kirkland, D. E. 1986. Corrugated Steel Plate Structures with
Continuous Longitudinal Stiffeners: Live Load Research and Recommended
Design Features for Short-Span Bridges. Transportation Research Record 1087,
TRB, Washington D. C., pp. 25-31.

176
Bakht, B. 1981. Soil-steel structure response to live loads. Journal of the Geotechnical
Engineering Division, ASCE, Vol. 107, No. GT6, pp. 779-798.

Bakht, B. 1985. Live Load Response of a Soil-Steel Structure with a Relieving Slab.
Transportation Research Record 1008, TRB, Washington D. C., pp. 1-7.

Bakht, B., and Agarwal, A. C. 1988. On Distress in Pipe-Arches. Canadian Journal of


Civil Engineering, Vol. 15, No. 4, pp. 589-595.

Beal, D. B. 1982. Field Tests of Long-Span Aluminum Culvert. Journal of the


Geotechnical Engineering Division, Vol. 108, No. GT6, pp. 873-890.

Booy, C. 1957. Flexible Conduit Studies. Prairie Farm Rehabilitation Administration,


Canada Department of Agriculture, Saskatoon, Saskatchewan, Canada.

Boulanger, R., Seed, R. B. and Schluter, J. C. 1989. Measurement and Analyses of


Deformed Flexible Box Culverts. Transportation Research Record 1231, TRB,
Washington D. C., pp. 25-35.

Burland, J. B. 1989. The Stiffness of Soils at Small Strains. Canadian Geotechnical


Journal, Vol. 26, No.4, pp. 499-516.

Burns, J. Q., and Richard, R. M. 1964. Attenuation of Stresses for Buried Cylinders.
Proceedings, Symposium on Soil Structure Interaction, University of Arizona,
Tucson, Arizona. pp. 378-392.

Burns, J. Q. 1965. An Analysis of Circular Cylindrical Shells Embedded in Elastic Media.


Ph.D. Dissertation, Engineering Mechanics Department, University of Arizona,
Tucson, Arizona.

Chang, C. S., Espinoza, J. M., and Selig, E. T. 1980. Computer Analysis of Newton
Creek Culvert. Journal of Geotechnical Engineering Division, ASCE, Vol. 106,
No GT5, pp. 531-556.

Chelliah, D. 1992. Experimental and Numerical Analysis of a Pipe Arch Culvert


Subjected to Exceptional Live Load. M.S. Thesis, Ohio University, Athens.

Cook, R. D., Malkus, D. S. and Plesha, M. E. 1989. Concepts and Applications of Finite
Element Analysis, Third Edition, New York: John Wiley & Sons, Inc.

Dar, S. M., and Bates, R. C.. 1974. Stress Analysis of Hollow Cylindrical Inclusions.
Journal of the Geotechnical Engineering Division, ASCE, Vol. 100, No. GT2.,
pp. 123-138.

Das, B. M. 2002. Principles of Geotechnical Engineering, 5th edition, Books/Cole.

177
Duncan, J. M. 2004. Friction Angles for Sand, Gravel and Rock. Notes of a lecture
presented at the Kenneth L. Lee Memorial Seminar Long Beach, California.

Duncan, J. M. 1979. Behavior and Design of Long-Span Metal Culvert. Journal of


Geotechnical Engineer Division, ASCE, Vol. 105, No GT3, pp. 399-418.

Duncan, J. M., Byrne, P., Wong, K. S., and Mabry, P. 1980. Strength, Stress-Strain and
Bulk Modulus Parameters for Finite Element Analyses of Stresses and
Movements in Soil Masses. Report No. UCB/GT/80-01, College of
Engineering, Office of Research Services, University of California, Berkeley,
California.

Duncan, J. M., and Jeyapalan, J. K. 1982. Deflection of Flexible Culverts due to Backfill
Compaction. Transportation Research Record 878, TRB, Washington D. C., pp.
10-17.

Duncan, J. M., and Drawsky, R. H. 1983. Design procedures for flexible metal culvert
structures. Report No. UCB/GT/83-04, University of California, Berkeley.

Duncan, J. M., Seed, R. B., and Drawsky, R. H. 1985. Design of Corrugated Metal Box
Culverts. Transportation Research Record 1008, TRB, Washington D. C., pp.
33-41.

Einstein, H. H., and Schwartz, C. W. 1979. Simplified Analysis for Tunnel Supports.
Journal of Geotechnical Engineering Division, Vol. 105, No. GT4, pp. 499-518.

Ghobrial, M., and Abdel-Sayed, G. 1985. Inelastic Buckling of Soil-Steel Structures.


Transportation Research Record 1008, TRB, Washington D. C., pp. 7-14.

Gorman, C. D. 1981. A Comparison of Field Test Results with Theoretical Analysis of


Lane Metal Products’ Low Profile Box Culvert. Report No. 79-67-1, Technical
Services Division, Eng. Dept. Bethlehem Steel Corp.

Hoeg, K. 1968. Stresses Against Underground Structural Cylinders. Journal of the Soil
Mechanics and Foundations Divison, ASCE, Vol. 94, SM4, pp. 833-858.

Huang, J. 2007. Evaluating Bridge Performance: Load Rating Bridges without Plans and
Experimental Displacement Influence Lines. Ph.D. Dissertation. University of
Delaware.

Hurd, J.O., and Sargand, S. 1988. Field Performance of Corrugated Metal Box Culverts.
Transportation Research Record 1191, TRB, Washington D. C., pp. 39-45.

Jardine, R.J., Potts, D. M., Fourie, A. B., and Burland, J. B. 1986. Studies of the
Influence of Non-linear Stress Strain Characteristics in Soil Structure
Interaction. Geotechnique, Vol. 36, No.3, pp. 377-396.

178
Katona, M. G., Smith, J. M., Odello, R. S., and Allgood, J. R. 1976. CANDE – A Modern
Approach for the Structural Design and Analysis of Buried Culverts. Report No.
FHWA-RD-77-5, U.S. Naval Civil Engineering Lab., Port Hueneme, CA.

Katona, M. G. 1978. Analysis of Long-Span Culverts by the Finite Element Method.


Transportation Research Record 678, TRB, Washington D. C., pp. 59-66.

Katona, M. G., Meinhert, D. F., Orillac, R., and Lee, C. H. 1979. Structural Evaluation of
New Concepts for Long-Span Culverts and Culvert Installations. Report No.
FHWA-RD-79-115, Federal Highway Administration, Washington, D. C.

Kay, J. N., Avalle, D. K., Flint, R. C. L., and Fitzhardinge, C. F. R. 1980. Instrumentation
of a Corrugated Steel-Soil Arch Overpass at Leigh Creek, South Australia.
Proceedings, 10th Conference of Australian Road Research Board, Vol. 10, No.
3, pp. 57-70.

Kay, J. N., and Flint, R. C. L. 1982. Heavy-Vehicle Loading of Arch Structures of


Corrugated Metal and Soil. Transportation Research Record 878, TRB,
Washington, D.C., pp. 34-37.

Kerh, T., and Yee, T. C. 2000. Analysis of a Deformed Three-Dimensional Culvert


Structure using Neural Networks. Journal of Advances in Engineering Software,
Vol. 31, pp. 367-375.

Kunecki, B., and Ebeltoft, R. 2007. Full-Scale Test of Steel Pipe-Arch Culvert and Finite
Element Analysis by Two Commercial Codes. Proceedings, TRB 2007 Annual
Meeting, pp. 1-11.

Lambe, T. W., and Whitman, R. V. 1969. Soil Mechanics, John Wiles & Sons.

Lee, K. L., Seed, H. B., and Dunlop, P. 1969. Effect of Transient Loading on the Strength
of Sand. Committee of VII ICSMFE, Proceedings of 7th International
Conference on Soil Mechanic and Foundation Engineering, London, Scotland:
Butterworths, pp. 239–247.

Luscher, U., and Hoeg, K. 1964. The Beneficial Action of the Surrounding Soil on the
Load Carrying Capacity of Buried Tubes. Proceedings of the Symposium on
Soil-Structure Interaction, University of Arizona, Tucson, Arizona., pp. 393-
402.

Manko, Z. Z, and Beben, D. 2005. Static Load Tests of a Road Bridge with a Flexible
Structure Made from Super Cor Type Steel Corrugated Plates. Journal of
Bridge Engineering, ASCE, Vol. 10, pp. 604-621.

179
Marston, A. 1930. The Theory of External Loads on Closed Conduits in the Light of the
Latest Experiments. Bulletin No. 96, Iowa, Engineering Experiment Station,
Ames, Iowa.

Masada, T. 2000. Modified Iowa Formula for Vertical Deflection of Buried Flexible Pipe.
Journal of Transportation Engineering, ASCE, Vol. 126, pp. 440-446.

Meyerhoff, G. G., and Baikie, L. D. 1963. Strength of Steel Culvert Sheets Bearing
against Compacted Sand Backfill. Highway Research Record No. 30, Highway
Research Board, National Academy of Sciences, Washington, D. C., pp. 1-19.

McVay, M. C., and Selig, E. T. 1982. Performance and Analysis of a Long-Span Culvert.
Transportation Research Record 878, TRB, Washington, D.C., pp. 23-29.

Moore, I. D. 1987. .Design Procedure for Calculating Elastic Buckling Strength of Metal
Culverts. Department of Civil Engineering and Surveying, University of
Newcastle., New South Wales, Australia, 1987.

Moore, I. O. 1989. Elastic Buckling of Buried Flexible Tubes - A Review of Theory and
Experiment. Journal of Geotechnical Engineering, American Society of Civil
Engineers, Vol. 115, No.3, pp. 340-358.

Moore, I. D., Selig E. T., and Haggag, A. 1988. Elastic Buckling Strength of Buried
Flexible Culverts. Transportation Research Record 1191, TRB, Washington,
D.C., pp. 57-64.

Moore, I. D., and Brachman, R. W. 1994. Three-Dimensional Analysis of Flexible


Circular Culverts. Journal of Geotechnical Engineering, ASCE, Vol. 120, No.
10, pp. 1829-1844.

Moore, I. D., and Taleb, B. 1999. Metal Culvert Response to Live Loading: Performance
of Three-Dimensional Analysis. Transportation Research Record 1656, TRB,
Washington, D.C., pp. 37-44.

Musser, S. C. 1989. CANDE-89 User Manual. Report No. FHWA-RD-89-169, Federal


Highway Administration.

NCSPA, 1995. Load Rating and Structural Evaluation of In-Service, Corrugated Steel
Structures. Design sheet No. 19. National Corrugated Steel Pipe Association,
Washington, D.C., pp. 1-12.

NCHRP 254, 1998. Service Life of Drainage Pipe” National Cooperative Highway
research Program. Transportation Research Board, National Academy Press,
Washington, D.C.

180
Ohio Department of Transportation. 1951-1989. Construction and Material Specifications.
Ohio Department of Transportation, Columbus, Ohio.

Ohio Department of Transportation. 2006. Manual of Bridge Inspection. Ohio


Department of Transportation, Columbus, Ohio.

Potter, J. C., Lewandowsky, L., and White Jr., D. W. 1991. Durability of Special
Coatings for Corrugated Steep Pipe. FHWA-91-FLP-006, U.S. Army
Waterways Experiment Station, Geotechnical Laboratory, Vicksburg, Missouri.

Rauch, A. F. 1990. Experimental and Numerical Investigation of a Deep-Corrugated


Steel, Box-Type Culvert. M.S. Thesis, Ohio University, Athens, Ohio.

Sargand, S., Masada, T., and Moreland, A. 2008. Measured Field Performance and
Computer Analysis of Large-Diameter Multiplate Steel Pipe Culvert Installed
in Ohio. Journal of Performance of Constructed Facilities, Vol. 22, No. 6,
ASCE, pp. 391-397.

Seed, R. B., and Ou, C. 1986. Measurements and Analyses of Compaction Effects on a
Long-Span Culvert. Transportation Research Record, No. 1087, TRB, National
Research Board Council, Washington, D.C., pp. 37-45.

Selig, E. T. 1975. Instrumentation of Large Buried Culverts. Performance Monitoring for


Geotechnical Construction, ASTM STP 584, American Society for Testing and
Materials, pp. 159-181.

Selig, E. T., and Calabrese, S. J. 1975. Performance of a Large Corrugated Steel Culvert.
Transportation Research Record 548, Soil and Rock Mechanics, Culverts, and
Compaction, TRB, Washington, D.C., pp. 62-76.

Selig, E. T, Lockhart, C. W., and Lautensleger, R. W. 1979. Measured Performance of


Newtown Creek Culvert. Journal of the Geotechnical Engineering Division,
ASCE, Vol. 105, No. GT9, pp. 1067-1087.

Selig, E. T., and Musser, S. C. 1985. Performance Evaluation of a Rib-Reinforced


Culvert. Transportation Research Record 1008, TRB, Washington, D.C., pp.
117-122.

Seed, R. B., and Ou, C. 1986. Measurements and Analyses of Compaction Effects on a
Long-Span Culvert. Transportation Research Record 1087, TRB, Washington,
D.C., pp. 37-45.

Spangler, M. G. 1941. The Structural Design of Flexible Pipe Culverts. Iowa State
College Engineering Experimental Station, Bulletin No. 153.

Spangler, M. G., and Handy, R. H. 1973. Soil Engineering. Harpercollins College Div.

181
Stavros A. J. 1984. Galvalume Corrugated Steel Pipe: A Performance Summary
Transportation Research Record 1001, TRB, National Research Board Council,
Washington, D.C., pp. 69-76.

Timoshenko, S. P., and Goodier, J. N. 1970. Theory of Elasticity. Third Edition, McGraw
Hill.

Townsend, H. E., and Zoccola, J. C. 1979. Atmospheric Corrosion Resistance of 55% Al-
Zn-Coated Sheet Steel: Thirteen-Year Test Results. Materials Performance,
Vol. 18, No. 10, pp. 13-20.

Trautmann, C. H., and Kulhawy, F. H. 1987. CUFAD―A Computer Program for


Compress and Uplift Foundation Analysis and Design. Report EL-4540-CMM,
Vol. 16, Electrical Power and Research Institute.

Vaslestad, J. 1989. Long-Term Behavior of Flexible Large-Span Culverts.


Transportation Research Record No. 1231, TRB, National Research Board
Council, Washington, D.C., pp. 14-24.

Watkins, R. K, and Anderson, L. R. 1999. Structural Mechanics of Buried Pipes. CRC


Press.

Webb, M. C., Selig, E. T., Sussmann, J. A., and McGrath, T. J. 1999. Field Tests of a
Large-Span Metal Culvert. Transportation Research Record 1656, TRB,
National Research Board Council, Washington, D.C., pp. 14-24.

Webb, M. C. 1999. Improved Design and Construction of Large-Span Culverts. Ph.D.


Dissertation, University of Massachusetts, Amherst.

White, H L, and Layer, J. P. 1960. The Corrugated Metal Conduit as a Compression Ring.
Highway Research Board Proceedings, Highway Research Board, Vol. 39, pp.
389-397.

White, K.R., Minor, J., and Derucher, K. N. 1992. Bridge Maintenance Inspection and
Evaluation (Second Edition, Revised and Expanded). Marcel Dekker.

Wong, K. S., and Duncan, J. M. 1974. Hyperbolic Stress-Strain Parameters for Nonlinear
Finite Element Analyses of Stresses and Movements in Soil Masses. Report No.
TE-74-3, College of Engineering, Office of Research Services, University of
California, Berkeley, California.

Wood, D. M. 2004. Geotechnical modeling. 1st edition, Spon Press.

182
APPENDIX A

EXPERIMENTAL TEST RESULTS FOR DEFLECTIONS

183
0.3
LC 9 10 mph 20 mph
CH #03 LC 8 5 mph
0.010
LC 10 30 mph
0.2
Deflection (mm)

Deflection (in.)
LC 3
LC 2 LC 7
LC 4 0.005
0.1 LC 1
LC 6
0 0
LC 5
15 mph
-0.1
0 200 400 600 800 1000 1200 1400
Time (s)
0.3
CH #02 0.010
CH #04
0.2
Deflection (mm)

Deflection (in.)
0.005
0.1

0 0

-0.1
0 200 400 600 800 1000 1200 1400
Time (s)
0.3
CH #01 0.010
0.2 CH #05
Deflection (mm)

Deflection (in.)

0.005
0.1

0 0

-0.1
0 200 400 600 800 1000 1200 1400
Time (s)

Figure A.1: Measured deflections in culvert ADA-041-0096 (1 mph = 1.6093 km/h)

184
0.5
CH #03 5 mph 15 mph
0.4 20 mph
LC 9 10 mph 30 mph 0.015
Deflection (mm)

40 mph

Deflection (in.)
0.3 LC 8
0.01
LC 3
0.2
LC 2 LC 4 LC 10
0.005
0.1 LC 1
LC 7
0 0
LC 5 LC 6
-0.1
0 200 400 600 800 1000 1200
Time (s)
0.5
CH #02
0.4 CH #04 0.015
Deflection (mm)

Deflection (in.)
0.3
0.01
0.2
0.005
0.1

0 0

-0.1
0 200 400 600 800 1000 1200
Time (s)
0.5
CH #01
0.4 0.015
CH #05
Deflection (mm)

Deflection (in.)

0.3
0.01
0.2
0.005
0.1

0 0

-0.1
0 200 400 600 800 1000 1200
Time (s)

Figure A.2: Measured deflections in culvert ALL-309-1664

185
0.8
0.03
CH #03 LC 10 15 mph
5 mph
30 mph
0.6 40 mph
Deflection (mm)

Deflection (in.)
LC 8 0.02
0.4
LC 3 LC 4 LC 9
0.01
0.2 LC 7
LC 2

0 0
LC 1 LC 5 LC 6 10 mph 20 mph
0 200 400 600 800 1000 1200
Time (s)
0.8
0.03
CH #02
0.6 CH #04
Deflection (mm)

Deflection (in.)
0.02
0.4

0.01
0.2

0 0

0 200 400 600 800 1000 1200


Time (s)
0.8
0.03
CH #01
0.6 CH #05
Deflection (mm)

Deflection (in.)

0.02
0.4

0.01
0.2

0 0

0 200 400 600 800 1000 1200


Time (s)

Figure A.3: Measured deflections in culvert ALL-696-0410

186
0.8 LC 9 0.03
5 mph 10 mph 20 mph
CH #03 LC 8
0.6
Deflection (mm)

40 mph 0.02

Deflection (in.)
0.4 LC 7
LC 3 LC 10 0.01
0.2 LC 2
LC 4
LC 1
0 LC 5 LC 6 0
15 mph 30 mph
-0.2
-0.01
0 500 1000 1500
Time (s)
0.8 0.03
CH #02
0.6 CH #04
Deflection (mm)

0.02

Deflection (in.)
0.4
0.01
0.2

0 0

-0.2
-0.01
0 500 1000 1500
Time (s)
0.8 0.03
CH #01
0.6 CH #05
Deflection (mm)

0.02
Deflection (in.)

0.4
0.01
0.2

0 0

-0.2
-0.01
0 500 1000 1500
Time (s)

Figure A.4: Measured deflections in culvert ATB-193-2316

187
0.6
5 mph 30 mph
0.5 CH #03 10 mph 15 mph 20 mph 0.02
LC 9
Deflection (mm)

LC 8

Deflection (in.)
0.4 40 mph 0.015
0.3
0.01
0.2 LC 3
LC 4
LC 2 LC 10 0.005
0.1 LC 7

0 LC 6 0
LC 1 LC 5
-0.1
0 200 400 600 800 1000 1200 1400
Time (s)
0.6
0.5 CH #02 0.020
CH #04
Deflection (mm)

Deflection (in.)
0.4 0.015
0.3
0.010
0.2
0.1 0.005

0 0
-0.1
0 200 400 600 800 1000 1200 1400
Time (s)
0.6
0.5 CH #01 0.020
CH #05
Deflection (mm)

Deflection (in.)

0.4 0.015
0.3
0.010
0.2
0.1 0.005

0 0
-0.1
0 200 400 600 800 1000 1200 1400
Time (s)

Figure A.5: Measured deflections in culvert ATB-307-1801

188
0.04
0.03 CH #03
0.001
Deflection (mm)

Deflection (in.)
0.02
0.01
0 0
-0.01
-0.02
-0.001
-0.03
0 100 200 300 400 500 600
Time (s)
0.04
0.03 CH #02
CH #04 0.001
Deflection (mm)

Deflection (in.)
0.02
0.01
0 0
-0.01
-0.02
-0.001
-0.03
0 100 200 300 400 500 600
Time (s)
0.04
0.03 CH #01
CH #05 0.001
Deflection (mm)

Deflection (in.)

0.02
0.01
0 0
-0.01
-0.02
-0.001
-0.03
0 100 200 300 400 500 600
Time (s)

Figure A.6: Measured deflections in culvert CLI-068-0942

189
1.0
LC 9 10 mph 15 mph
CH #03 LC 8 30 mph
0.8 40 mph
0.03
Deflection (mm)

0.6

Deflection (in.)
LC 3 LC 10 0.02
0.4 LC 4
LC 2 0.01
0.2 5 mph
LC 7
0 LC 5 0
LC 1 LC 6
-0.2 20 mph
-0.01
-0.4
0 200 400 600 800 1000
Time (s)
1.0
0.8 CH #02
0.03
CH #04
Deflection (mm)

0.6

Deflection (in.)
0.02
0.4
0.2 0.01

0 0
-0.2 -0.01
-0.4
0 200 400 600 800 1000
Time (s)
1.0
0.8 CH #01 0.03
CH #05
Deflection (mm)

0.6
Deflection (in.)

0.02
0.4
0.2 0.01

0 0
-0.2 -0.01
-0.4
0 200 400 600 800 1000
Time (s)

Figure A.7: Measured deflections in culvert COL-170-1981

190
0.8 LC 9 5 mph 15 mph 20 mph 30 mph
CH #03 0.03
40 mph
Deflection (mm)

0.6

Deflection (in.)
LC 8 0.02
LC 3 LC 4
0.4
LC 7
LC 10
0.2 LC 2
0.01

LC 6
0 0
LC 1 LC 5 10 mph
-0.2
0 200 400 600 800
Time (s)

0.8 0.03
CH #02
CH #04
Deflection (mm)

0.6

Deflection (in.)
0.02
0.4
0.01
0.2

0 0

-0.2
0 200 400 600 800
Time (s)

0.8 CH #01 0.03


CH #05
Deflection (mm)

0.6
Deflection (in.)

0.02
0.4
0.01
0.2

0 0

-0.2
0 200 400 600 800
Time (s)

Figure A.8: Measured deflections in culvert COL-172-1209

191
0.5
CH #03 5 mph 15 mph 20 mph
0.4 LC 9 0.015
LC 8 30 mph
Deflection (mm)

Deflection (in.)
40 mph
0.3
LC 7 0.010
LC 10
0.2 LC 3 LC 4
LC 2 0.005
0.1
LC 1 LC 6
0 0
LC 5 10 mph
-0.1
0 200 400 600 800 1000 1200 1400
Time (s)
0.5
CH #02
0.4 0.015
CH #04
Deflection (mm)

Deflection (in.)
0.3
0.010
0.2
0.005
0.1

0 0

-0.1
0 200 400 600 800 1000 1200 1400
Time (s)
0.5
CH #01
0.4 CH #05 0.015
Deflection (mm)

Deflection (in.)

0.3
0.010
0.2
0.005
0.1

0 0

-0.1
0 200 400 600 800 1000 1200 1400
Time (s)

Figure A.9: Measured deflections in culvert DEF-018-1560

192
0.10
CH #03 20 mph
0.08 LC 8 10 mph 0.003
Deflection (mm)

Deflection (in.)
LC 7 15 mph 30 mph
0.06 LC 9 5 mph
LC 6 40 mph
LC 3 0.002
0.04
LC 2 LC 10
0.02 0.001
LC 1

0 0
LC 4 LC 5
-0.02
0 500 1000 1500
Time (s)
0.10
0.08 CH #02
CH #04 0.003
Deflection (mm)

Deflection (in.)
0.06
0.04 0.002
0.02
0.001
0
-0.02 0

-0.04
0 500 1000 1500
Time (s)
0.10
CH #01
0.08 0.003
Deflection (mm)

Deflection (in.)

0.06
0.002
0.04
0.001
0.02

0 0

-0.02
0 500 1000 1500
Time (s)

Figure A.10: Measured deflections in culvert DEL-037-0333

193
1.5
LC 8 LC 9 10 mph 15 mph 0.05
1.2 CH #03
0.04
Deflection (mm)

30 mph

Deflection (in.)
0.9 LC 7
LC 3 0.03
0.6 LC 6 LC 10
LC 2 0.02
40 mph
0.3 0.01
LC 4 5 mph
LC 5
0 0
LC 1
-0.3 20 mph -0.01
-0.02
-0.6
0 200 400 600 800 1000 1200 1400 1600
Time (s)
1.5
1.2 CH #02 0.05
CH #04 0.04
Deflection (mm)

Deflection (in.)
0.9
0.03
0.6
0.02
0.3 0.01
0 0
-0.3 -0.01
-0.02
-0.6
0 200 400 600 800 1000 1200 1400 1600
Time (s)
1.5
1.2 CH #05 0.05
0.04
Deflection (mm)

Deflection (in.)

0.9
0.03
0.6
0.02
0.3 0.01
0 0
-0.3 -0.01
-0.02
-0.6
0 200 400 600 800 1000 1200 1400 1600
Time (s)

Figure A.11: Measured deflections in culvert FAY-734-0790

194
0.6
LC 8 5 mph 15 mph
0.5 CH #03 30 mph 0.020
40 mph
LC 9
Deflection (mm)

Deflection (in.)
0.4 LC 7 0.015
0.3
0.010
0.2 LC 3 LC 10
LC 2 LC 6
0.1 0.005

0 0
LC 1 LC 4 LC 5 10 mph 20 mph
-0.1
0 200 400 600 800 1000 1200 1400 1600
Time (s)
0.6
0.5 CH #02 0.020
CH #04
Deflection (mm)

Deflection (in.)
0.4 0.015
0.3
0.010
0.2
0.1 0.005

0 0
-0.1
0 200 400 600 800 1000 1200 1400 1600
Time (s)
0.6
0.5 CH #01 0.020
CH #05
Deflection (mm)

Deflection (in.)

0.4 0.015
0.3
0.010
0.2
0.1 0.005

0 0
-0.1
0 200 400 600 800 1000 1200 1400 1600
Time (s)

Figure A.12: Measured deflections in culvert FRA-062-2688

195
1.2
1 CH #03 LC 10 5 mph 10 mph 0.04
20 mph
Deflection (mm)

30 mph

Deflection (in.)
0.8 40 mph 0.03
LC 9
0.6
LC 4 LC 5 0.02
0.4
LC 3 LC 8 0.01
0.2 LC 2
0 LC 7 0
LC 6
LC 1 15 mph
-0.2
0 200 400 600 800 1000 1200
Time (s)
1.2
1.0 CH #02 0.04
CH #04
Deflection (mm)

Deflection (in.)
0.8 0.03
0.6
0.02
0.4
0.2 0.01

0 0
-0.2
0 200 400 600 800 1000 1200
Time (s)
1.2
1.0 CH #01 0.04
CH #05
Deflection (mm)

Deflection (in.)

0.8 0.03
0.6
0.02
0.4
0.2 0.01

0 0
-0.2
0 200 400 600 800 1000 1200
Time (s)

Figure A.13: Measured deflections in culvert GAL-218-0578

196
1.2
LC 8 10 mph 20 mph
1 30 mph
0.04
CH #03 40 mph
Deflection (mm)

LC 9

Deflection (in.)
0.8 0.03
LC 3
0.6
0.02
0.4 LC 4
LC 2 LC 10
0.2 0.01
LC 7
0 0
LC 1 LC 5 LC 6 5 mph 15 mph
-0.2
0 200 400 600 800 1000
Time (s)
1.2
1 CH #02 0.04
CH #04
Deflection (mm)

Deflection (in.)
0.8 0.03
0.6
0.02
0.4
0.2 0.01

0 0
-0.2
0 200 400 600 800 1000
Time (s)
1.2
1 CH #01 0.04
CH #05
Deflection (mm)

Deflection (in.)

0.8 0.03
0.6
0.02
0.4
0.2 0.01

0 0
-0.2
0 200 400 600 800 1000
Time (s)

Figure A.14: Measured deflections in culvert HIG-134-0782

197
1.0
10 mph 15 mph 20 mph
0.8 CH #03 LC 10
LC 9 0.03
30 mph 40 mph
Deflection (mm)

Deflection (in.)
0.6
LC 8 0.02
LC 4
0.4
LC 3
LC 5 0.01
0.2 LC 2
5 mph
LC 1
0 0
LC 6 LC 7
-0.2
0 300 600 900 1200 1500 1800 2100
Time (s)
1.0

0.8 CH #02
CH #04 0.03
Deflection (mm)

Deflection (in.)
0.6
0.02
0.4

0.2 0.01

0 0
-0.2
0 300 600 900 1200 1500 1800 2100
Time (s)
1.0

0.8 CH #01
CH #05 0.03
Deflection (mm)

Deflection (in.)

0.6
0.02
0.4

0.2 0.01

0 0
-0.2
0 300 600 900 1200 1500 1800 2100
Time (s)

Figure A.15: Measured deflections in culvert HOC-093-1975

198
1.2
10 mph
1.0 CH #03 LC 8 LC 9 20 mph 30 mph 40 mph 0.04
Deflection (mm)

Deflection (in.)
0.8
LC 7 0.03
0.6
LC 3
0.02
0.4 LC 2 LC 10
LC 4
0.2 0.01
5 mph
0 LC 1 LC 5 LC 6
15 mph
0
-0.2
0 500 1000 1500 2000
Time (s)
1.2
1.0 CH #02 0.04
CH #04
Deflection (mm)

Deflection (in.)
0.8 0.03
0.6
0.02
0.4
0.2 0.01

0 0
-0.2
0 500 1000 1500 2000
Time (s)
1.2
1.0 CH #01 0.04
CH #05
Deflection (mm)

Deflection (in.)

0.8 0.03
0.6
0.02
0.4
0.2 0.01

0 0
-0.2
0 500 1000 1500 2000
Time (s)

Figure A.16: Measured deflections in culvert HOC-664-1172

199
0.5
CH #03 LC 10 5 mph 15 mph 30 mph
0.4 0.015
LC 9 40 mph
Deflection (mm)

Deflection (in.)
0.3 LC 8
0.010
0.2 LC 4 LC 7
LC 2 0.005
0.1
LC 5 LC 6 10 mph 20 mph
0 0
LC 1 LC 3
-0.1
0 500 1000 1500
Time (s)
0.5
CH #02
0.4 0.015
CH #04
Deflection (mm)

Deflection (in.)
0.3
0.010
0.2
0.005
0.1

0 0

-0.1
0 500 1000 1500
Time (s)
0.5
CH #01
0.4 0.015
Deflection (mm)

Deflection (in.)

0.3
0.010
0.2
0.005
0.1

0 0

-0.1
0 500 1000 1500
Time (s)

Figure A.17: Measured deflections in culvert KNO-013-2162

200
0.03
0.001
0.02 CH #03
Deflection (mm)

Deflection (in.)
0.01

0 0
-0.01

-0.02
-0.001
-0.03
0 50 100 150 200 250 300 350
Time (s)
0.03
0.001
0.02 CH #02
CH #04
Deflection (mm)

Deflection (in.)
0.01

0 0

-0.01

-0.02
-0.001
-0.03
0 50 100 150 200 250 300 350
Time (s)
0.03
0.001
0.02 CH #05
Deflection (mm)

Deflection (in.)

0.01

0 0

-0.01

-0.02
-0.001
-0.03
0 50 100 150 200 250 300 350
Time (s)

Figure A.18: Measured deflections in culvert KNO-062-1480

201
1.5 LC 9 0.06
5 mph 10 mph 20 mph 30 mph
1.2 LC 8 40 mph
Deflection (mm)

0.04

Deflection (in.)
0.9
LC 3 LC 10
0.6
LC 2 LC 4 0.02
0.3
0 0
LC 7
-0.3 LC 1 15 mph
-0.6 LC 5 CH #03 -0.02
LC 6
0 200 400 600 800
Time (s)

1.5 0.06
CH #04
1.2
Deflection (mm)

0.04

Deflection (in.)
0.9
0.6
0.02
0.3
0 0
-0.3
-0.6 -0.02

0 200 400 600 800


Time (s)

1.5 0.06
CH #01
1.2
CH #05 0.04
Deflection (mm)

Deflection (in.)

0.9
0.6 0.02
0.3
0 0
-0.3
-0.6 -0.02

0 200 400 600 800


Time (s)

Figure A.19: Measured deflections in culvert LAK-608-0606

202
0.3
CH #03 10 mph 30 mph 0.010
LC 8 LC 9 15 mph
0.2 20 mph
Deflection (mm)

Deflection (in.)
LC 7
40 mph
LC 6 LC 10 0.005
0.1 LC 3
LC 2
5 mph
0 0
LC 4 LC 5
LC 1
-0.1
0 200 400 600 800 1000 1200 1400
Time (s)
0.3
CH #02 0.010
0.2 CH #04
Deflection (mm)

Deflection (in.)
0.005
0.1

0 0

-0.1
0 200 400 600 800 1000 1200 1400
Time (s)
0.3
CH #05 0.01
0.2
Deflection (mm)

Deflection (in.)

0.005
0.1

0 0

-0.1
0 200 400 600 800 1000 1200 1400
Time (s)

Figure A.20: Measured deflections in culvert LIC-079-2276

203
0.5
LC 8 LC 9
0.4 CH #03
0.015
Deflection (mm)

Deflection (in.)
0.3 LC 4 LC 7 LC 10
LC 3
0.010
0.2 LC 2
0.1 LC 6 5 mph 0.005
LC 5
10 mph
0 15 mph 30 mph 0
LC 1
20 mph 40 mph
-0.1
0 200 400 600 800 1000 1200 1400 1600
Time (s)
0.5

0.4 CH #02
CH #04 0.015
Deflection (mm)

Deflection (in.)
0.3
0.010
0.2
0.005
0.1

0 0

-0.1
0 200 400 600 800 1000 1200 1400 1600
Time (s)
0.5

0.4 0.015
Deflection (mm)

Deflection (in.)

0.3
0.010
0.2 No data available due to noise
0.005
0.1

0 0

-0.1
0 200 400 600 800 1000 1200 1400 1600
Time (s)

Figure A.21: Measured deflections in culvert LIC-661-0805

204
0.3
5 mph 10 mph 20 mph 40 mph 0.010
CH #03 LC 9
LC 8
0.2
Deflection (mm)

Deflection (in.)
LC 4 LC 7 LC 10
LC 6 0.005
0.1 LC 3
LC 2

0 LC 5 0
15 mph 30 mph
LC 1
-0.1
0 200 400 600 800 1000 1200
Time (s)
0.3
CH #02 0.010
0.2 CH #04
Deflection (mm)

Deflection (in.)
0.005
0.1

0 0

-0.1
0 200 400 600 800 1000 1200
Time (s)
0.3
CH #01 0.010
0.2
Deflection (mm)

Deflection (in.)

0.005
0.1

0 0

-0.1
0 200 400 600 800 1000 1200
Time (s)

Figure A.22: Measured deflections in culvert MAD-056-1490

205
0.8
LC 8 10 mph 20 mph 30 mph
0.03
CH #03 LC 7
LC 9
0.6 5 mph
Deflection (mm)

Deflection (in.)
LC 6 0.02
0.4
LC 2
LC 3 LC 10
0.2 0.01

LC 4 LC 5
0 LC 1
0
40 mph
15 mph
-0.2
0 500 1000 1500
Time (s)
0.8
0.03
CH #02
0.6 CH #04
Deflection (mm)

Deflection (in.)
0.02
0.4

0.2 0.01

0 0

-0.2
0 500 1000 1500
Time (s)
0.8
0.03
CH #01
0.6 CH #06
Deflection (mm)

Deflection (in.)

0.02
0.4

0.2 0.01

0 0

-0.2
0 500 1000 1500
Time (s)

Figure A.23: Measured deflections in culvert MAD-665-1071

206
0.5
LC 8 5 mph 15 mph 30 mph
0.4 CH #03 LC 9 40 mph
LC 7
0.015
Deflection (mm)

Deflection (in.)
0.3 LC 10
LC 6
0.010
0.2 LC 3
LC 2
0.005
0.1 LC 4 LC 5
LC 1
0 0
10 mph 20 mph
-0.1
0 200 400 600 800 1000 1200 1400
Time (s)
0.5

0.4 CH #02
CH #04 0.015
Deflection (mm)

Deflection (in.)
0.3
0.010
0.2
0.005
0.1

0 0

-0.1
0 200 400 600 800 1000 1200 1400
Time (s)
0.5

0.4 CH #01
CH #05 0.015
Deflection (mm)

Deflection (in.)

0.3
0.010
0.2
0.005
0.1

0 0

-0.1
0 200 400 600 800 1000 1200 1400
Time (s)

Figure A.24: Measured deflections in culvert MAR-229-0093

207
0.5
40 mph
0.4 CH #03 LC 8 5 mph
20 mph 0.015
Deflection (mm)

LC 9 10 mph

Deflection (in.)
0.3 LC 7 30 mph
LC 3
LC 6 0.010
0.2 LC 2
LC 10

0.1 0.005
LC 4 LC 5
0 0
LC 1 15 mph
-0.1
0 200 400 600 800 1000
Time (s)
0.5

0.4 CH #02
CH #04 0.015
Deflection (mm)

Deflection (in.)
0.3
0.010
0.2

0.1 0.005

0 0
-0.1
0 200 400 600 800 1000
Time (s)
0.5

0.4 CH #01
CH #05 0.015
Deflection (mm)

Deflection (in.)

0.3
0.010
0.2
0.005
0.1

0 0

-0.1
0 200 400 600 800 1000
Time (s)

Figure A.25: Measured deflections in culvert MAR-309-1248

208
0.3
LC 8 LC 10 15 mph CH #03 0.010
LC 9
5 mph 20 mph 30 mph
0.2
Deflection (mm)

Deflection (in.)
LC 7 40 mph
LC 6 0.005
0.1 LC 3

0 LC 5 0
LC 2
LC 4 10 mph
LC 1
-0.1
0 500 1000 1500
Time (s)
0.3
CH #02 0.010
0.2 CH #04
Deflection (mm)

Deflection (in.)
0.005
0.1

0 0

-0.1
0 500 1000 1500
Time (s)
0.3
CH #05 0.01
0.2
Deflection (mm)

Deflection (in.)

0.005
0.1

0 0

-0.1
0 500 1000 1500
Time (s)

Figure A.26: Measured deflections in culvert MAR-423-0214

209
0.6
LC 8 LC 9 5 mph 10 mph 20 mph 30 mph
0.5 CH #03 0.020
15 mph 40 mph
0.4
Deflection (mm)

0.015

Deflection (in.)
0.3 LC 7 LC 10
LC 3 LC 4 0.010
0.2
LC 2
0.1 0.005
0 0
LC 1 LC 5 LC 6
-0.1
-0.005
-0.2
0 200 400 600 800 1000
Time (s)
0.6
0.5 CH #02 0.020
0.4 CH #04
Deflection (mm)

Deflection (in.)
0.015
0.3
0.010
0.2
0.1 0.005
0 0
-0.1
-0.005
-0.2
0 200 400 600 800 1000
Time (s)
0.6
0.5 CH #05 0.020
0.4
Deflection (mm)

Deflection (in.)

0.015
0.3
0.010
0.2
0.1 0.005
0 0
-0.1
-0.005
-0.2
0 200 400 600 800 1000
Time (s)

Figure A.27: Measured deflections in culvert MAR-423-0979

210
0.5
LC 8 LC 9 15 mph 20 mph 30 mph
0.4 10 mph
LC 10 40 mph 0.015
Deflection (mm)

Deflection (in.)
LC 4 5 mph
0.3
LC 3
0.010
0.2
LC 7
0.1 0.005

0 LC 5 LC 6 0
LC 1 LC 2 CH #03
-0.1
0 200 400 600 800 1000 1200 1400 1600
Time (s)
0.5

0.4 CH #02
0.015
Deflection (mm)

CH #04

Deflection (in.)
0.3
0.010
0.2

0.1 0.005

0 0
-0.1
0 200 400 600 800 1000 1200 1400 1600
Time (s)
0.5

0.4 CH #01
0.015
Deflection (mm)

Deflection (in.)

0.3
0.010
0.2

0.1 0.005

0 0
-0.1
0 200 400 600 800 1000 1200 1400 1600
Time (s)

Figure A.28: Measured deflections in culvert MRW-095-1637

211
0.3
CH #03 10 mph 15 mph 0.010
LC 9 20 mph
LC 8
0.2
Deflection (mm)

Deflection (in.)
LC 7
LC 4
LC 6 LC 10 30 mph 40 mph
LC 3 0.005
0.1
LC 2
5 mph
0 0
LC 5
LC 1
-0.1
0 200 400 600 800 1000 1200
Time (s)
0.3
CH #02 0.010
CH #04
0.2
Deflection (mm)

Deflection (in.)
0.005
0.1

0 0

-0.1
0 200 400 600 800 1000 1200
Time (s)
0.3
CH #01 0.010
0.2 CH #05
Deflection (mm)

Deflection (in.)

0.005
0.1

0 0

-0.1
0 200 400 600 800 1000 1200
Time (s)

Figure A.29: Measured deflections in culvert MRW-095-1796

212
1.2
5 mph 15 mph 20 mph
1.0 CH #03 LC 8 LC 9 0.04
0.8 30 mph 40 mph
Deflection (mm)

Deflection (in.)
0.03
0.6 LC 5
LC 2 0.02
0.4
LC 3 0.01
0.2 LC 10
0 0
LC 7
-0.2 LC 1 LC 4 LC 6 10 mph
-0.01
-0.4
0 300 600 900 1200 1500 1800
Time (s)
1.2
1.0 CH #02 0.04
0.8 CH #04
Deflection (mm)

Deflection (in.)
0.03
0.6
0.02
0.4
0.2 0.01
0 0
-0.2 -0.01
-0.4
0 500 1000 1500
Time (s)
1.2
1.0 CH #01 0.04
0.8 CH #05
Deflection (mm)

Deflection (in.)

0.03
0.6
0.02
0.4
0.2 0.01
0 0
-0.2 -0.01
-0.4
0 500 1000 1500
Time (s)

Figure A.30: Measured deflections in culvert NOB-083-0091

213
0.4 0.015
LC 9 10 mph 20 mph
0.3 CH #03 40 mph
LC 8
0.010
Deflection (mm)

Deflection (in.)
LC 10
0.2 LC 4 LC 7
LC 3 0.005
0.1
LC 2
LC 5 LC 6
0 0
5 mph
LC 1
-0.1 30 mph
15 mph -0.005
-0.2
0 500 1000 1500 2000
Time (s)
0.4 0.015
0.3 CH #02
CH #04 0.010
Deflection (mm)

Deflection (in.)
0.2
0.005
0.1

0 0

-0.1
-0.005
-0.2
0 500 1000 1500 2000
Time (s)
0.4 0.015
CH #01
0.3
CH #05 0.010
Deflection (mm)

Deflection (in.)

0.2
0.005
0.1

0 0

-0.1
-0.005
-0.2
0 500 1000 1500 2000
Time (s)

Figure A.31: Measured deflections in culvert PIC-138-0091

214
0.15
5 mph 10 mph
LC 8 20 mph 30 mph 0.005
CH #03
LC 9 40 mph
0.10 LC 7 0.004
Deflection (mm)

LC 4

Deflection (in.)
LC 3
0.003
LC 2
0.05 0.002
LC 6 0.001
0 LC 10 0
LC 1 -0.001
LC 5 15 mph
-0.05
0 100 200 300 400 500 600 700 800
Time (s)
0.15
CH #02 0.005
0.10 CH #04 0.004
Deflection (mm)

Deflection (in.)
0.003
0.05 0.002
0.001
0 0
-0.001
-0.05
0 100 200 300 400 500 600 700 800
Time (s)
0.15
CH #05 0.005
0.10 0.004
Deflection (mm)

Deflection (in.)

0.003
0.05 0.002
0.001
0 0
-0.001
-0.05
0 100 200 300 400 500 600 700 800
Time (s)

Figure A.32: Measured deflections in culvert PIC-752-0748

215
2.5
CH #03 LC 8 LC 9 5 mph 15 mph 20 mph
0.08
2.0 30 mph
Deflection (mm)

Deflection (in.)
1.5 LC 3 0.06
LC 10
1.0 LC 4 0.04
LC 7
0.5 LC 2 0.02
LC 1 40 mph
0 10 mph
0
LC 5 LC 6
-0.5 -0.02
-1.0
0 300 600 900 1200 1500 1800
Time (s)
2.5
2.0 CH #02 0.08
CH #04
Deflection (mm)

Deflection (in.)
1.5 0.06
1.0 0.04
0.5 0.02
0 0
-0.5 -0.02
-1.0
0 300 600 900 1200 1500 1800
Time (s)
2.5
2.0 CH #01 0.08
Deflection (mm)

Deflection (in.)

1.5 0.06
1.0 0.04
0.5 0.02
0 0
-0.5 -0.02
-1.0
0 300 600 900 1200 1500 1800
Time (s)

Figure A.34: Measured deflections in culvert SAN-412-0575

216
1.4 5 mph 15 mph
CH #03 LC 8 LC 9 10 mph 20 mph 0.05
1.2
30 mph 40 mph
Deflection (mm)

Deflection (in.)
1.0 0.04
LC 3
0.8 LC 4 LC 10
0.03
0.6 LC 2 LC 5 0.02
0.4
LC 1 0.01
0.2 LC 7
0 LC 6
0
-0.2
0 200 400 600 800 1000 1200 1400 1600
Time (s)

1.4
1.2 CH #02 0.05
CH #04
Deflection (mm)

Deflection (in.)
1.0 0.04
0.8 0.03
0.6
0.02
0.4
0.2 0.01
0 0
-0.2
0 200 400 600 800 1000 1200 1400 1600
Time (s)

1.4
CH #01 0.05
1.2 CH #05
Deflection (mm)

Deflection (in.)

1.0 0.04
0.8 0.03
0.6
0.02
0.4
0.2 0.01
0 0
-0.2
0 200 400 600 800 1000 1200 1400 1600
Time (s)

Figure A.35: Measured deflections in culvert SAN-510-0412

217
0.6
10 mph
0.5 CH #03 LC 8 5 mph 15 mph 0.020
20 mph 30 mph
LC 9
Deflection (mm)

40 mph

Deflection (in.)
0.4 0.015
0.3 LC 6 LC 7
LC 2 LC 3 0.010
0.2 LC 10
0.1 0.005
LC 5
LC 1 LC 4
0 0
-0.1
0 300 600 900
Time (s)
0.6
0.5 CH #02 0.020
CH #04
Deflection (mm)

Deflection (in.)
0.4 0.015
0.3
0.010
0.2
0.1 0.005

0 0
-0.1
0 300 600 900
Time (s)
0.6
0.5 CH #01 0.020
CH #05
Deflection (mm)

Deflection (in.)

0.4 0.015
0.3
0.010
0.2
0.1 0.005

0 0
-0.1
0 300 600 900
Time (s)

Figure A.36: Measured deflections in culvert TRU-193-0215

218
APPENDIX B

EXPERIMENTAL TEST RESULTS FOR STRAINS

219
80 CH #07 CH #08 CH #09
40
0
-40
-80
-120
80
40
0
-40
-80
-120 CH #10 CH #12

80
40
0
-40
-80
-120 CH #13 CH #15
0 200 400 600 800 1000
80
40
Strain (µs)

0
-40
-80 CH #18
-120 CH #16

80
40
0
-40
-80 CH #14
-120 CH #17

80
40 CH #11 CH #19 CH #20
0
-40
-80
-120
0 200 400 600 800 1000
Time (s)

Figure B.1: Measured strain-time histories for culvert MAD-038-0935

220
5 CH #07 CH #08 CH #09
0
-5
-10

5 CH #10 CH #12
0
-5
-10

5 CH #13
0
-5
-10

5
Strain (µs)

0
-5 CH #16
-10 CH #18

5
0
-5 CH #14
-10 CH #17

5
0
-5
-10 CH #11 CH #19 CH #20

0 200 400 600 800 1000 1200


Time (s)

Figure B.2: Measured strain-time histories for culvert MRW-095-1796

221
80
40
0
-40
-80 CH #08 CH #09

80 CH #10 CH #12
40
0
-40
-80

80
40 CH #13
0
-40
-80

80
CH #16
Strain (µs)

40 No data available
CH #18
0
-40
-80

80
40 CH #14
0
-40
-80

80
40
0
-40 CH #19
-80 CH #20

0 200 400 600 800 1000


Time (s)

Figure B.3: Measured strain-time histories for culvert JAC-327-0077

222
10
0 No data available due to noise
-10
-20

10 CH #12

0
-10
-20

10 CH #13 CH #15
0
-10
-20

10 CH #16 CH #18
Strain (µs)

0
-10
-20

10
0
-10
-20 CH #14 CH #17

10 CH #11 CH #19 CH #20


0
-10
-20
0 200 400 600 800 1000 1200
Time (s)

Figure B.4: Measured strain-time histories for culvert ADA-041-0096

223
10 CH #07 CH #08 CH #09
0
-10
-20

10
0
-10
CH #10
-20 CH #12

10
CH #13 CH #15
0
-10
-20

10
CH #16 CH #18
Strain (µs)

0
-10
-20

10
0
-10
CH #14
-20 CH #17

10
0
-10 CH #11
CH #19
-20 CH #20
0 200 400 600 800 1000 1200
Time (s)

Figure B.5: Measured strain-time histories for culvert ALL-309-1664

224
40 CH #08 CH #09
20
0
-20
-40

40 CH #10 CH #12
20
0
-20
-40
40
CH #13
20 CH #15
0
-20
-40
40
CH #18
Strain (µs)

20 CH #16
No data available
0
-20
-40
40
CH #14
20 CH #17
0
-20
-40
40
CH #11 CH #19 CH #20
20
0
-20
-40
0 200 400 600 800 1000 1200
Time (s)

Figure B.6: Measured strain-time histories for culvert ALL-696-0410

225
10
0
-10 CH #07 No data available
-20 CH #08
-30 CH #09

10
0
-10
-20
-30 CH #10

10
0
-10
-20
-30 CH #13 CH #15

10 CH #16 CH #18
Strain (µs)

0
-10
-20
-30

10
0
-10
-20 CH #14
-30 CH #17

10
0
-10
CH #11
-20 CH #20
-30 CH #19
0 200 400 600 800 1000
Time (s)

Figure B.7: Measured strain-time histories for culvert ATB-193-2316

226
10
0
-10 CH #07
CH #08
-20 CH #09

10
0
-10
-20 CH #10

10
0
-10
CH #13
-20 CH #15

10
Strain (µs)

0
-10
CH #16
-20 CH #18

10
0
-10
CH #14
-20 CH #17

10
CH #11
0 CH #20 No data available
CH #19
-10
-20
0 200 400 600 800 1000 1200
Time (s)

Figure B.8: Measured strain-time histories for culvert ATB-307-1801

227
15 CH #08 CH #09
0
-15
-30

15
CH #12
0
-15
-30
Sheet1

15
CH #13
0 CH #15
No dada available

-15
-30

15
Strain (µs)

0
-15
-30 CH #16

15
0
-15
CH #14
-30 CH #17

15
0
-15
CH #11
-30 CH #20

0 200 400 600 800 1000 1200


Time (s)

Figure B.9: Measured strain-time histories for culvert COL-170-1981

228
15 CH #07 CH #08 CH #09
0
-15
-30

15
0
-15
CH #10
-30 CH #12

15
CH #15
0
-15
-30

15
Strain (µs)

0
-15
CH #16
-30 CH #18

15
0
-15
CH #14
-30 CH #17

15
0
-15 CH #11
CH #20
-30 CH #19
0 200 400 600 800
Time (s)

Figure B.10: Measured strain-time histories for culvert COL-172-1209

229
10
0
-10
-20
CH#07 CH#08 CH#09
20
10
0
-10
-20 CH#10 CH#12
-30

10 CH#15
0
-10
-20

10 CH#18
Strain (µs)

0
-10
-20

10
0
-10
-20 CH#14 CH#17

10 CH#11 CH#20
0
-10
-20
0 200 400 600 800 1000 1200 1400
Time (s)

Figure B.11: Measured strain-time histories for culvert DEF-018-1560

230
5
0
-5
-10 CH #07
-15 CH #08

5
0
-5
-10 CH #10
-15 CH #12

5
0
-5
-10 CH #13
-15 CH #15

5
Strain (µs)

0
-5
-10 CH #16
-15 CH #18

5
0
-5
-10 CH #14
-15 CH #17

5
0
-5
-10 CH #11
-15 CH #19
0 300 600 900 1200
Time (s)

Figure B.12: Measured strain-time histories for culvert DEL-037-0333

231
45 CH #07
30 CH #09
15
0
-15
-30

45 CH #10
30 CH #12
15
0
-15
-30

45 CH #13
30 CH #15
15
0
-15
-30

45 CH #16
Strain (µs)

30 CH #18
15
0
-15
-30

45 CH #14
30 CH #17
15
0
-15
-30

45 CH #19
30 CH #20
15
0
-15
-30
0 500 1000 1500 2000
Time (s)

Figure B.13: Measured strain-time histories for culvert FAY-734-0790

232
10 CH #07 CH #08 CH #09
0
-10
-20
-30

10
0
-10
-20
CH #10
-30 CH #12

10
0
-10
-20
CH #13
-30 CH #15

10
Strain (µs)

0
-10
-20 CH #16
-30 CH #18

10
0
-10
-20 CH #14
-30 CH #17

10
0
-10
-20
CH #11
-30 CH #19
0 200 400 600 800 1000
Time (s)

Figure B.14: Measured strain-time histories for culvert FRA-062-2688

233
30 CH #07 CH #08 CH #09
15
0
-15
-30

30 CH #10 CH #12
15
0
-15
-30

30
CH #13
15 CH #15
0
-15
-30

30
CH #16
Strain (µs)

15 CH #18
0
-15
-30

30
15
0
-15 CH #14
-30 CH #17

30
CH #11 CH #19 CH #20
15
0
-15
-30
0 200 400 600 800 1000
Time (s)

Figure B.15: Measured strain-time histories for culvert GAL-218-0578

234
60
30
0
-30
-60 CH #07 CH #09

60
CH #10 CH #12
30
0
-30
-60

60
CH #13 CH #15
30
0
-30
-60
60
CH #16
Strain (µs)

30
0
-30
-60

60
CH #14 CH #17
30
0
-30
-60

60
CH #11 CH #19 CH #20
30
0
-30
-60
0 200 400 600 800 1000 1200
Time (s)

Figure B.16: Measured strain-time histories for culvert HIG-134-0782

235
25 CH #08 CH #09
0
-25
-50

25 CH #10 CH #12
0
-25
-50

25 CH #13
0
-25
-50

25
Strain (µs)

0
-25 CH #16
-50 CH #18

25 CH #14
0
-25
-50

25
0
-25
-50 CH #11 CH #19 CH #20
0 500 1000 1500 2000
Time (s)

Figure B.17: Measured strain-time histories for culvert HOC-093-1975

236
20
0
-20 CH #07
CH #08
-40
CH #09

20
0
-20
CH #10
-40 CH #12

20 CH #13 CH #15
0
-20
-40

20 CH #16 CH #18
Strain (µs)

0
-20
-40

20
0
-20 CH #14
-40 CH #17

20
0
-20
-40 CH #11 CH #19
0 500 1000 1500 2000
Time (s)

Figure B.18: Measured strain-time histories for culvert HOC-664-1172

237
5
0
-5
-10 CH #08
-15 CH #09

5
0
-5
-10 CH #10
-15 CH #12

5
0
-5
-10 CH #13
-15 CH #15

5
Strain (µs)

0
-5
-10 CH #16
-15 CH #18

5
0
-5
-10 CH #14
-15 CH #17

5
0
-5 CH #11
-10 CH #19
-15 CH #20
0 200 400 600 800 1000
Time (s)

Figure B.19: Measured strain-time histories for culvert KNO-013-2162

238
5
0
-5
-10
-15 CH #07
-20 CH #09
-25
5
0
-5
-10
-15 CH #13
-20 CH #15
-25
5
0
-5
-10
-15 CH #16
-20 CH #18
-25
5
0
Strain (µs)

-5
-10
-15 CH #14
-20 CH #17
-25
5
0
-5
-10
-15
-20 CH #11 CH #19 CH #20
-25
5
0
-5
-10
-15
-20 CH #21 CH #22 CH #23
-25
0 500 1000 1500 2000
Time (s)

Figure B.20: Measured strain-time histories for culvert LIC-079-2276

239
10 CH #07 CH #08 CH #09
0
-10
-20
-30
10
0
-10
-20 CH #10
-30 CH #12

10
0
-10
-20 CH #13
-30 CH #15

10
Strain (µs)

0
-10
-20 CH #16
-30 CH #18

10
0
-10
-20 CH #14
-30 CH #17

10
0
-10
CH #11
-20 CH #19
-30 CH #20
0 200 400 600 800 1000 1200 1400 1600
Time (s)

Figure B.21: Measured strain-time histories for culvert LIC-661-0805

240
15
0
-15
-30
-45 CH #07 CH #08 CH #09

15
0 CH #10
-15 CH #12
-30
-45

15
0
-15
-30 CH #13
-45 CH #15

15
Strain (µs)

0
-15
-30 CH #16
-45 CH #18

15
0
-15
-30 CH #14
-45 CH #17

15
0
-15
-30 CH #11
-45 CH #19
0 500 1000 1500
Time (s)

Figure B.22: Measured strain-time histories for culvert MAD-665-1071

241
10 CH #07 CH #08 CH #09
0
-10
-20

10
0
-10
CH #10
-20 CH #12

10
0
-10
CH #13
-20 CH #15

10
Strain (µs)

0
-10
CH #16
-20 CH #18

10
0
-10
CH #14
-20 CH #17

10
0
-10
-20 CH #11 CH #19 CH #20
0 200 400 600 800 1000 1200
Time (s)

Figure B.23: Measured strain-time histories for culvert MAR-229-0093

242
5
0
-5
-10
-15 CH #08 CH #09

5
0
-5
-10 CH #10
-15 CH #12

5
CH #13 CH #15
0
-5
-10
-15

5
Strain (µs)

0
-5
-10
-15 CH #16

5
0
-5
-10 CH #14
-15 CH #17

5
0
-5
-10
-15 CH #11 CH #19 CH #20
0 100 200 300 400 500 600 700 800
Time (s)

Figure B.24: Measured strain-time histories for culvert MAR-309-1248

243
5 CH #07 CH #08 CH #09
0
-5
-10

5
0
-5 CH #10
-10 CH #12

5
0
-5 CH #13
-10 CH #15

5
Strain (µs)

0
-5
CH #16
-10 CH #18

5 CH #14
CH #17 No data available
0
-5
-10

5
0
-5
-10 CH #11 CH #19 CH #20
0 200 400 600 800 1000 1200 1400 1600
Time (s)

Figure B.25: Measured strain-time histories for culvert MAR-423-0214

244
CH #07 CH #08 CH #09
10
0
-10
-20

10 CH #10
0
-10
-20

10 CH #13 CH #15
0
-10
-20

10
Strain (µs)

0
-10
CH #16
-20 CH #18

10
0
-10
CH #14
-20 CH #17

10
0
-10
-20 CH #11 CH #19 CH #20
0 200 400 600 800
Time (s)

Figure B.26: Measured strain-time histories for culvert MAR-423-0379

245
10
CH #07 CH #08 CH #09
0
-10
-20

10
0
-10 CH #10
CH #12
-20
10
0
-10 CH #13
CH #15
-20
10
Strain (µs)

0
-10 CH #16
CH #18
-20
10
0
-10 CH #14
CH #17
-20
10
0
-10
CH #11 CH #19 CH #20
-20
0 200 400 600 800 1000 1200
Time (s)

Figure B.27: Measured strain-time histories for culvert MRW-019-0456

246
15
0
-15
-30 CH #09

15
0
-15
-30
CH #10 CH #12
15
CH #13 CH #15
0
-15
-30

15
CH #16 CH #18
Strain (µs)

0
-15
-30

15
0
-15
-30 CH #14

15
0
-15
CH #19
-30 CH #20
0 200 400 600 800 1000 1200 1400 1600
Time (s)

Figure B.28: Measured strain-time histories for culvert MRW-095-1637

247
20
0
-20
-40
-60 CH #07 CH #08 CH #09

20
0
-20
-40 CH #10
-60 CH #12

20
0
-20
-40 CH #13
-60 CH #15

20
Strain (µs)

0
-20
-40 CH #16
-60 CH #18

20
0
-20
-40 CH #14
-60 CH #17
0 500 1000 1500

20
0
-20
-40 CH #11
-60 CH #19
0 500 1000 1500
Time (s)

Figure B.29: Measured strain-time histories for culvert NOB-083-0419

248
10
5 CH #07 CH #08 CH #09
0
-5
-10
-15
-20

10
5
0
-5
-10
-15 CH #10 CH #12
-20

10
5
0
-5
-10 CH #13
-15 CH #15
-20

10
5 CH #16 CH #18
Strain (µs)

0
-5
-10
-15
-20

10
5 CH #21 CH #22
0
-5
-10
-15
-20

10
5
0
-5
-10
-15
-20 CH #23 CH #24 CH #25
0 500 1000 1500 2000 2500
Time (s)

Figure B.30: Measured strain-time histories for culvert PIC-138-0091

249
4
0
-4
CH #07
-8 CH #08
-12 CH #09

4
0
-4
-8 CH #10
-12 CH #12

4
0
-4
-8 CH #13
-12 CH #15

4
Strain (µs)

0
-4
-8 CH #16
-12 CH #18

4
0
-4
-8 CH #14
-12 CH #17

4
0
-4
CH #11
-8 CH #19
-12 CH #20
0 100 200 300 400 500 600 700
Time (s)

Figure B.31: Measured strain-time histories for culvert PIC-752-0748

250
50
0
-50
-100 CH #07 CH #08 CH #09

50
0
-50
-100 CH #10 CH #12

50 CH #13
0
-50
-100

50 CH #16
Strain (µs)

0
-50
-100

50 CH #14 CH #17
0
-50
-100

50 CH #11 CH #19 CH #20


0
-50
-100
0 500 1000 1500
Time (s)

Figure B.32: Measured strain-time histories for culvert SAN-412-0575

251
40
CH #08
20 CH #09
0
-20
-40

40
CH #10
20 CH #12
0
-20
-40

40
CH #13
20 CH #15
No data available
0
-20
-40

40
CH #16
Strain (µs)

20
0
-20
-40

40
CH #14
20
0
-20
-40

40
CH #11 CH #19 CH #20
20
0
-20
-40
0 200 400 600 800 1000 1200 1400 1600
Time (s)

Figure B.33: Measured strain-time histories for culvert SAN-510-0412

252
20 CH #07 CH #08 CH #09

0
-20
-40

20
CH #10
0
-20
-40

20
0
-20
CH #13
-40 CH #15

20
Strain (µs)

0
-20
CH #16
-40 CH #18

20
0
-20
CH #14
-40 CH #17

20
0
-20
CH #11
-40 CH #19
0 100 200 300 400 500 600 700 800
Time (s)

Figure B.34: Measured strain-time histories for culvert TRU-193-0215

253
APPENDIX C

DUNCAN’S HYPERBOLIC SOIL MODEL

254
C.1 Introduction
The finite element program, CANDE introduced in Chapter 6 is one of the
powerful programs that can be used to analyze stresses and movements in soil masses and
various soil-structure interaction problems. The main issue in solving soil problems is the
representation of soil stress-strain characteristics in a reasonable way. Duncan (1980) soil
model provides hyperbolic stress-strain relations in nonlinear incremental analyses of soil
deformation. The model parameters can be obtained from conventional triaxial
compression tests. The model is relatively simple to implement in finite element codes.
Therefore, these factors have extensively been used in CANDE.

C.2 Hyperbolic stress-strain relationships


The stress-strain behavior of soil is assumed to be governed by the generalized
Hooke’s Law for elastic materials. The relationship is expressed in Equation for
conditions of plane strain.
 ∆σ x  1 −ν v 0   ∆ε x 
  E    ∆ε 
 ∆σ y  =  v 1 −ν 0  y  (C.1)
  (1 + ν )(1 − 2ν )  0 0 2(1 − 2ν )  ∆γ xy 
∆τ xy    
where ∆σ x and ∆σ y are normal stress increments, ∆τ xy is shear stress increment, ∆ε x

and ∆ε y is normal strain increments, ∆γ xy is shear strain increment, E is Young’s

modulus, and ν is Poisson’s ratio. Generally, the terms of E and ν are changed to E
and bulk modulus ( B ) in soil mechanics as shown in Equation C.2. The bulk modulus is
defined as the pressure increase needed to cause a specified relative decrease in volume.
E
B= (C.2)
3(1 − 2ν )
From Equations C.1 and C.2, the following equation can be derived.
 ∆σ x  3 B + E 3 B − E 0   ∆ε x 
  3B   
 ∆σ y  = 3B − E 3B + E 0   ∆ε y  (C.3)
  9B − E  0 0 E   ∆γ xy 
∆τ xy    

255
C.3 Theoretical development
The hyperbola shown in Figure C.1 can be represented by the following equation
suggested by Konder et al. (1963).
ε
(σ 1 − σ 3 ) = (C.4)
1 ε
+
Ei (σ 1 − σ 3 )ult

where σ 1 is the major principal stress, σ 3 is the minor principal stress, Ei is the initial

elastic modulus and ε is the compressive axial strain. The ultimate stress difference
(σ 1 − σ 3 )ult is reached only at infinite strains. The failure stress will be reached before the

ultimate stress. Thus, the failure ratio ( R f ) is always less than 1.0. For most soils, R f

varies from 0.5 to 0.9.


(σ 1 − σ 3 ) f
Rf = (C.5)
(σ 1 − σ 3 )ult
If the confining pressure is increased, the stress-strain curve will be steeper. Thus, the
initial elastic modulus and ultimate pressure increase as confining pressure increases. The
variation of Ei and σ 3 is represented by the following equation, which is suggested by
Janbu (1963).
n
σ 
Ei = K ⋅ Pa  3  (C.6)
 Pa 
where Pa is atmospheric pressure (14.7 psi = 101.35 kPa), K and n can be obtained
from triaxial tests with more than three data points. At least more than three triaxial tests
need to be conducted to obtain the parameters. A plot of Equation C.6 on log-log paper
yields a straight line of the form.
E  σ 
log10  i  = log10 K + n log10  3  (C.7)
 Pa   Pa 
σ  E
If σ 3 equals Pa , that is, log10  3  = 0 , K = i , we can obtain K value. Also, the
 P3  Pa

parameter n is equal to the slope of this line on a log-log scale.

256
ACTUAL
ε
(σ 1 − σ 3 ) =
1 ε
+
Ei (σ 1 − σ 3 )ult

TRANSFORMED
ε
(σ 1 − σ 3 ) =
1 ε
+
Ei (σ 1 − σ 3 )ult

Figure C.1: Hyperbolic representation of a strain curve (Duncan, et al., 1980)

In order to determine the angle of internal friction ( φ ) and the cohesion ( c ), the
Mohr-Coulomb failure criterion (Figure C.2) is applied. Generally, for a cohesionless
material such as clean sand, c can be taken as zero. For a cohesive soil such as clay, c is
will be different from zero.
The variation of the bulk modulus with the minor principal stress is characterized
with Equation C.8.
m
σ 
B = K b Pa  3  (C.8)
 pa 

257
Figure C.2: Mohr-Coulomb failure criterion.

In order to determine Kb and m , a linear regression needs to be performed to determine


the best straight line fit on log-log paper with the following relationship.
 B σ 
log10   = log10 K b + m log10  3  (C.9)
 pa   pa 
The values of K b and m are the intercept and slope, respectively.

258
APPENDIX D

CONDITION ASSESSMENT GUIDELINE FOR CORRUGATED


METAL CULVERTS

(ODOT MANUAL OF BRIDGE INSPECTION, 2006)

259
D.1 Introduction
Appendix D presents ODOT’s assessment method for the bridge inspection
condition. The condition coding system used for the General Appraisal was developed by
the Federal Highway Administration. The State of Ohio has used its own code prior to the
Federal code and the Ohio code is still used in the Bridge Inspection Report (BR-86).
However, the Federal code will be introduced only in this study because most of the
recent bridge conditions are recorded officially using the Federal code. The following
Federal codes are used to summarize the condition of the General Appraisal. In this study,
the Federal code is applied to culvert type bridges which convey water or passageway
through an embankment.
General Appraisal Number Bridge condition
9 Excellent condition
8 Very good condition
7 Good condition
6 Satisfactory condition
5 Fair condition
4 Poor condition
3 Serious condition
2 Critical condition
1 "Imminent" failure condition
0 Failed condition

D.2 Condition assessment


Culverts are designed to support dead loads, like soil and pavement, imposed on
the culvert, and live loads passing over the culvert. For the assessment of culvert
condition, culvert inspection includes deterioration, settlement, open joints, plugging,
deflection, cracks or signs of movement. The following General Appraisal Numbers
(GANs) for corrugated metal culvert are used by ODOT to reflect the culvert condition.

260
GAN Condition for corrugated metal culverts except pipe-arch culverts

9 New condition
8 Good condition
Shape: good, smooth curvature in barrel; span dimension within 10
percent of design
Seams and Joints: tight, no openings
Metal: superficial rust, no pitting
7 Generally good condition
Shape: generally good, top half of pipe smooth but minor flattening of
bottom; span dimension within 10 percent of design
Seams or Joints: minor joint or seam openings, potential for backfill
infiltration
Metal: moderate rust, slight pitting
6 Fair condition
Shape: fair, top half has smooth curvature but bottom half has flattened
significantly, span dimension within 10 percent of design.
Seams or joints: minor cracking at bolts is prevalent in one seam in lower
half of pipe. Evidence of backfill infiltration through seams or joints
Metal: fairly heavy rust, moderate pitting
5 Generally fair condition
Shape: generally fair, significant distortion at isolated locations in top half
and extreme flattening of invert, span dimension within 10 to 15 percent
greater than design
Seams or joints: moderate cracking at bolt holes along one seam near
bottom of pipe, deflection of pipe caused by backfill infiltration through
seams or Joints
Metal: scattered heavy rust, deep pitting
4 Marginal condition
Shape: marginal significant distortion throughout length of pipe, lower
third may be kinked, span dimension within 10 percent to 15 percent
greater than design, noticeable dip in guardrail over pipe.
Seams or Joints: Moderate cracking at bolt holes on one seam near top of
pipe, deflection caused by loss of backfill through open joints.
Metal: extensive heavy rust, deep pitting, heavy loss of section, chipping
hammer could easily punch a hole thru metal.
3 Poor condition
Shape: poor with extreme deflection at isolated locations, flattening of
crown, crown radius 20 to 30 feet, span dimension in excess of 15 percent
greater than design
Seams or joints: 3 inch long cracks at bolt holes on one seam
Metal: extensive heavy rust, deep pitting, scattered perforations
2 Critical condition
Shape: critical, extreme distortion and deflection throughout pipe
flattening of crown, crown radius over 30 feet, span dimension more than
20 percent greater than design

261
Seams or joints: plate cracked from bolt to bolt on one seam
Metal: extensive perforations due to rust
1 Critical condition
Shape: partially collapsed with crown in reverse curve
Seams or joints: failed
Road: closed to traffic
0 Critical condition
Pipe: totally failed
Road: closed to traffic

GAN Condition for corrugated metal pipe arches

9 New condition
8 Good condition
Shape: good with smooth curvature; span dimension with less than 3
percent greater than design
Joints or seams: good condition
Metal: minor construction defects, protective coatings intact, superficial
rust, no pitting
7 Generally good condition
Shape: generally good, smooth curvature in top half, bottom flattened but
still curved; span dimension within 3 to 5 percent greater than design
Joints or seams: minor joint or seam openings, infiltration of backfill
possible
Metal: moderate rust, slight pitting
6 Fair condition
Shape: fair, smooth curvature in top half, bottom flat, span dimension
within 5 percent greater than design
Joints or seams: minor cracking all along one seam, minor joint openings
with evidence of infiltration
Metal: fairly heavy rust, moderate pitting
5 Generally fair condition
Shape: generally fair, significant distortion in top in one location; bottom
has slight reverse curvature in one location
Joints and seams: moderate cracking at bolt holes along a seam in one
section, backfill being lost through seam or joint causing slight deflection
Metal: scattered heavy rust, deep pitting
4 Marginal condition
Shape: marginal, signification distortion all along top of arch, bottom has
reverse curve; span dimension more than 7 percent greater than design,
noticeable dip in guardrail over pipe.
Joints and seams: moderate cracking all along one seam; backfill
infiltration causing major deflection
Metal: extensive heavy rust, deep pitting, heavy loss of section, chipping
hammer could easily punch a hole thru metal.

262
3 Poor condition
Shape: poor, extreme deflection in top arch in one section; bottom has
reverse curvature throughout; span dimension more than 7 percent greater
than design
Seams: seam cracked 3 in. on each side of bolt holes
Metal: extensive heavy rust, deep pitting, scattered perforations
2 Critical condition
Shape: critical, extreme deflection along top of pipe; span dimension
more than 7 percent greater than design
Seams or joints: seam cracked from bolt to bolt down one seam
Metal: extensive perforations due to rust
1 Critical condition
Shape: structure partially collapsed
Seams or joints: seam failed
Road: closed to traffic
0 Critical condition
Shape: structure collapsed
Road: closed to traffic

263

Das könnte Ihnen auch gefallen