160
Georgia Department of Transportation Office of Materials and Research GDOT Research Project No. 2002 Final Report LABORATORY AND 3D NUMERICAL MODELING WITH FIELD MONITORING OF REGIONAL BRIDGE SCOUR IN GEORGIA Submitted by Terry Sturm 1 , Fotis Sotiropoulos 1 , Mark Landers 2 , Tony Gotvald 2 , SeungOh Lee 1 , Liang Ge 1 , Ricardo Navarro 1 , and Cristian Escauriaza 1 1 School of Civil and Environmental Engineering Georgia Institute of Technology Atlanta, GA 30332 and 2 U. S. Geological Survey 3039 Amwiler Rd. Suite 130 Atlanta, GA 30360 August 2004

RP 2002 Final Report - Georgia

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Page 1: RP 2002 Final Report - Georgia

Georgia Department of Transportation Office of Materials and Research

GDOT Research Project No. 2002 Final Report

LABORATORY AND 3D NUMERICAL MODELING WITH FIELD MONITORING OF REGIONAL BRIDGE SCOUR IN GEORGIA

Submitted by

Terry Sturm1, Fotis Sotiropoulos1, Mark Landers2, Tony Gotvald2, SeungOh Lee1, Liang Ge1, Ricardo Navarro1, and Cristian Escauriaza1

1School of Civil and Environmental Engineering Georgia Institute of Technology

Atlanta, GA 30332

and

2U. S. Geological Survey 3039 Amwiler Rd. Suite 130

Atlanta, GA 30360

August 2004

Page 2: RP 2002 Final Report - Georgia

TECHNICAL REPORT STANDARD TITLE PAGE

1. Report No. FHWA-GA-04-2002

2. Government Accession No.

3. Recipient's Catalog No.

5. Report Date August 2004

4. Title and Subtitle Laboratory and 3D Numerical Modeling with Field Monitoring of Regional Bridge Scour in Georgia 6. Performing Organization Code

7. Author(s) Terry Sturm, Fotis Sotiropoulos, SeungOh Lee, Liang Ge, Ricardo Navarro, Cristian Escauriaza (Georgia Institute of Technology)

Mark Landers2, Tony Gotvald, (U.S. Geological Survey)

8. Performing Organ. Report No.: 2002

10. Work Unit No.

9. Performing Organization Name and Address School of Civil and Environmental Engineering; Georgia Institute of Technology; Atlanta, GA 30332 U. S. Geological Survey; 3039 Amwiler Rd. Suite 130; Atlanta, GA 30360

11. Contract or Grant No.

13. Type of Report and Period Covered Final; 2000-2004

12. Sponsoring Agency Name and Address Georgia Department of Transportation Office of Materials and Research 15 Kennedy Drive Forest Park, Georgia 30297-2534

14. Sponsoring Agency Code

15. Supplementary Notes Prepared in cooperation with the U.S. Department of Transportation Federal Highway Administration. 16. Abstract Field measurements, laboratory modeling, and 3D numerical modeling of scour around bridge foundations were combined and applied to specific bridges in Georgia in order to elucidate the physics of the scouring process and improve scour predictions. Field data were collected at four Georgia bridge sites using fixed instrumentation and mobile instrumentation. These data include continuous scour measurements at several locations around a single pier at each site and continuous velocity measurements at two sites. Two bridges were modeled in the laboratory, including the full river bathymetry. Also, sediments sampled at the foundations of 10 Georgia bridges were tested in the laboratory to obtain erodibility parameters. Finally, a comprehensive 3D numerical model was applied to the bridges in this study. The field results revealed several important aspects of bridge scour processes, including the dynamics of live-bed scour, simultaneous occurrence of contraction and pier scour, and cyclical scour and fill associated with the tidal cycle. The 3D model was validated by laboratory measurements and revealed details of the complex flow field around natural river bridge foundations. Comparisons of laboratory scour depths with existing scour formulas highlighted the difficulties in scaling of scour depths from laboratory to field; however, a successful modeling strategy was applied. The laboratory model not only reproduced measured maximum scour depths in the field for bank-full and extreme flood events, but also the details of cross-sectional changes immediately upstream of the bridge. The laboratory erosion tests illustrated the regional variability of erosion parameters and the variability associated with sediment stratification at a particular site. Erosion parameters were successfully correlated with some easily measured sediment properties. Continuing studies will focus on contraction scour and the development of a scour prediction methodology that incorporates the knowledge gained from combined field, laboratory, and numerical studies.

17. Key Words Bridge scour, numerical modeling, sediment properties

18. Distribution Statement

19. Security Classif. (of this report) Unclassified

20. Security Classif. (of this page) Unclassified

21. No. of Pages 158

22. Price

Form DOT 1700.7 (8-69) *A color edition of this report is available at http://topps.dot.state.ga.us/homeoffs/fpmr.www/internal/b-admin/research/r-rpts-online.shtml.

Page 3: RP 2002 Final Report - Georgia

i

ACKNOWLEDGMENTS

The authors gratefully acknowledge the expert guidance and assistance provided by David Jared,

Special Research Engineer, of the Georgia DOT Research Office throughout the project as well

as the support and valuable suggestions offered by Sam Teal of the Bridge Design Office. The

assistance of Tom Scruggs and Shannon Shaneyfelt of the Geotechnical Bureau in obtaining the

Shelby tube cores is greatly appreciated. The project was also made possible by the support of

Georgene Geary, State Materials and Research Engineer; Rick Deaver, Chief of Research and

Development; and Paul Liles, State Bridge Engineer.

Page 4: RP 2002 Final Report - Georgia

ii

EXECUTIVE SUMMARY

Field measurements, laboratory modeling, and 3D numerical modeling of scour around bridge foundations have been successfully combined and applied to specific bridges in Georgia in order to elucidate the physics of the scouring process and improve scour predictions. Field data have been collected at four bridge sites in Georgia using fixed instrumentation in combination with mobile instrumentation. These data include continuous scour measurements at several locations around a single pier at each site as well as continuous velocity measurements at two of the sites. Two of these bridges have been modeled in the laboratory including the full river bathymetry. In addition, sediments sampled at the foundations of 10 bridges in Georgia have been tested in the laboratory to obtain erodibility parameters. Finally, a comprehensive 3D numerical model, including a novel approach to grid generation for bridge foundations in natural rivers and a state-of-the-art turbulence submodel, have been applied to the bridges in this study. The field results have revealed several important aspects of bridge scour processes including the dynamics of live-bed scour, simultaneous occurrence of contraction and pier scour, and cyclical scour and fill associated with the tidal cycle. In addition, the field data have proved to be invaluable for comparison with laboratory model results and have validated the need for additional continuous and simultaneous measurements of scour depths and flow fields. The 3D numerical model has revealed details of the complex flow field around natural river bridge foundations including large-scale coherent vortex shedding in the vicinity of the bridge foundations with multiple vortices having axes parallel and perpendicular to the bed. The 3D model has been validated by laboratory measurements of both velocity profiles and turbulence characteristics in the vicinity of the bridge pier bents. In addition, the 3D model has been demonstrated to be a powerful tool for understanding not only the flow field but the coupling between the flow field and measured scour patterns. Comparisons of laboratory scour depths with existing scour formulas have highlighted some of the difficulties in scaling of scour depths from the laboratory to the field; however, a successful modeling strategy has been applied in which the model sediment size is selected to obtain approximate Froude number similarity in the clear-water scour regime while maintaining geometric similarity of depths and bridge dimensions at a reasonably large ratio of model pier size to sediment diameter. The laboratory model successfully reproduced not only measured maximum scour depths in the field for both bank-full and extreme flood events, but also the details of cross-sectional changes immediately upstream of the bridge. The laboratory erosion tests illustrated the regional variability of erosion parameters as well as the variability associated with sediment stratification at a particular site. Erosion parameters were successfully correlated with some easily measured sediment properties. These advances in field data collection, 3D numerical modeling, and laboratory modeling of bridge scour, as well as in measurement and prediction of sediment erodibility properties, point the way to improved scour prediction techniques to be finalized in Phase 2 of this research.

Page 5: RP 2002 Final Report - Georgia

iii

TABLE OF CONTENTS

Page No. ACKNOWLEDGMENTS i EXECUTIVE SUMMARY ii LIST OF FIGURES vi LIST OF TABLES x CHAPTER 1. INTRODUCTION 1

MOTIVATION 1 BACKGROUND 3 RESEARCH OBJECTIVES 8 ORGANIZATION OF THE REPORT 9

CHAPTER 2. FIELD DATA COLLECTION 10

FIXED FIELD INSTRUMENTATION 10 MOBILE FIELD INSTRUMENTATION 11 STREAMBED SEDIMENT SAMPLING 12 FOUR SELECTED SITES 12

Chattahoochee River near Cornelia, GA 12 Site Description 12 Fixed Field Instrumentation 14 Data Collected 14

Ocmulgee River at Macon, GA 17 Site Description 17 Fixed Field Instrumentation 20 Data Collected 20

Fint River at Bainbridge, GA 24 Site Description 24 Fixed Field Instrumentation 25 Data Collected 26

Darien River at Darien, GA 28 Site Description 28 Fixed Field Instrumentation 29 Data Collected 30

CHAPTER 3. LABORATORY MODEL STUDIES

INTRODUCTION 33 MODELING CONSIDERATIONS 33 EXPERIMENTAL METHODS 35

Model Construction 35 Experimental Instrumentation 41 Experimental Procedure 43

Page 6: RP 2002 Final Report - Georgia

iv

Page No.

EXPERIMENTAL RESULTS 43 Summary of Data 43 Approach Velocity and Turbulence 47 Near-Field Turbulence Intensity 48 Scour Contours and Near-Field Velocities 50 Comparison of Laboratory Scour Data with Prediction Formulas 54

CHAPTER 4. THREE-DIMENSIONAL NUMERICAL MODELING 59

INTRODUCTION 59 NUMERICAL METHOD 61 GRID GENERATION FOR COMPLEX BATHYMETRY AND OBSTRUCTIONS 61 APPLICATION OF THE 3D MODEL 64 CONCLUSIONS 73

CHAPTER 5. SEDIMENT EROSION PROPERTIES 74

INTRODUCTION 74 SEDIMENT EROSION RESISTANCE 74 EROSION MEASUREMENTS 79 EROSION RELATIONSHIPS 82 EXPERIMENTAL METHODS 82

Sample Locations and Properties 82 Laboratory Flume Measurements 85

RESULTS AND ANALYSIS 88 Observed Erosion Behavior 88 Experimental Results 88 Prediction of Critical Shear Stress and Erosion Rate Constant 92 Critical Shear Stress 93 Erosion Rate Constant 97 Discussion 100

CHAPTER 6. COMPARISONS OF NUMERICAL MODEL, LABORATORY, AND FIELD RESULTS 102

INTRODUCTION 102 NUMERICAL MODEL VALIDATION 103 Chattahoochee River near Cornelia 103 Flint River at Bainbridge 106 Ocmulgee River at Macon 109 FLOW STRUCTURES AND SCOUR 111 COMPARISONS OF LABORATORY AND FIELD RESULTS 116

Page 7: RP 2002 Final Report - Georgia

v

Page No. Chattahoochee River near Cornelia 116 Flint River at Bainbridge 120 SUMMARY 123

CHAPTER 7. SUMMARY AND CONCLUSIONS 125 REFERENCES 130 APPENDIX A. EROSION AND SOIL PROPERTY TEST RESULTS 135

Page 8: RP 2002 Final Report - Georgia

vi

LIST OF FIGURES

Figure Page No.

2.1. Chattahoochee River near Cornelia, GA. 13

2.2. Fathometer and velocity meter layout at Cornelia. 14

2.3 Fathometer data from Cornelia. 15

2.4 Cross-section comparison at the upstream side of the bridge at Cornelia. 16

2.5. Layout of surveyed cross-sections at Cornelia. 17

2.6. Particle size distribution of bed material samples at Cornelia. 18

2.7. Ocmulgee River at Macon, GA. 19

2.8. Fathometer layout for Ocmulgee River at Macon, GA. 20

2.9 Fathometer data from Macon. 21

2.10 Cross-section comparison at the upstream side of the bridge at Macon. 22

2.11. Layout of surveyed cross-sections at Macon. 23

2.12. Particle size distribution of bed material samples at Macon. 23

2.13. Flint River at Bainbridge, GA. 24

2.14. Fathometer and velocity meter layout at Bainbridge. 26

2.15. Cross-section comparison at Bainbridge, GA. 27

2.16. Layout of surveyed cross-sections at Bainbridge. 27

2.17. Particle size distribution of bed material samples at Bainbridge at left (L), right (R), and center (C) of main channel for cross section numbers shown in Fig. 2.16. 28

2.18. Darien River at Darien, GA. 29

2.19. Fathometer layout at Darien. 30

2.20. Fathometer data from Darien. 31

2.21. Fathometer data at downstream side of left bridge fender. 31

Page 9: RP 2002 Final Report - Georgia

vii

Figure Page No.

2.22. Layout of surveyed cross-sections at Darien. 32

2.23. Particle size distribution of bed material samples at Darien. 32

3.1. Sketch of central pier bent of Chattahoochee River bridge near Cornelia, GA with prototype elevations and dimensions. 37

3.2. Sketch of central pier bent of Flint River bridge at Bainbridge with prototype elevations and dimensions. 39

3.3. Model of Chattahoochee River bridge near Cornelia. 40

3.4. Model of Flint River bridge at Bainbridge. 40

3.5. Scour development with time for Chattahoochee River model. 46

3.6. Longitudinal and vertical relative turbulence intensity profiles at approach section for Chattahoochee River model. 48

3.7. Longitudinal and vertical relative turbulence intensity profiles at approach section for Flint River model. 48

3.8. Near-field turbulence intensities for Chattahoochee River model. 49

3.9. Shaded scour contours and velocity vectors after scour at 40 percent of the approach depth for bank-full flow in the Chattahoochee flat-bed model. 51

3.10. Shaded scour contours and velocity vectors after scour at 40 percent of the approach depth for bank-full flow in the Chattahoochee River model. 51

3.11. Scour contours and velocity vectors after scour for bank-full flow in Flint River flat-bed model. 53

3.12. Scour contours and velocity vectors after scour for flood flow in Flint River model. 55

3.13. Comparison of measured scour depths in Chattahoochee River flat-bed model with scour prediction formulas for bank-full flow. 57

3.14. Comparison of measured scour depths in Chattahoochee River model with scour prediction formulas for bank-full flow and flood flow. 57

3.15. Comparison of measured scour depths in Flint River flat-bed model with scour prediction formulas for bank-full flow. 59

Page 10: RP 2002 Final Report - Georgia

viii

Figure Page No.

3.16. Comparison of measured scour depths in Flint River model with scour prediction formulas for flood flow. 59

4.1. Numerical geometry of a single bent of piers on natural river reach. 62

4.2 Overset grid system. 62

4.3. Dimensionless velocity-time history at two different points A and B. 66

4.4. Visualization of instantaneous flow field in the vicinity of bridge piers. 67

4.5. Bridger piers mounted on flat bed. 69

4.6. Streamwise velocity contours. 70

4.7. Snapshot of instantaneous streamlines of the large-scale flow. 70

4.8. Instantaneous streamwise velocity contours. 72

4.9 Turbulence kinetic energy profile. 72

5.1. Shields diagram for direct determination of critical shear stress of coarse-grained sediments. 77

5.2. Shelby tube core sample locations. 84

5.3. Recirculating flume for erosion testing. 86

5.4. Erosion rate vs. applied shear stress relationships for sediment samples with erosion rates up to 1.1 kg/m2/s. 90

5.5. Erosion rate vs. applied shear stress relationships for sediment samples with erosion rates up to 0.06 kg/m2/s. 91

5.6. Comparison of the measured and predicted critical shear stress parameter using Eq. 5.6. 96

5.7 Comparison of measured data and calculated values using Eq. 5.6 plotted on Shields’ diagram format. 97

5.8. Comparison of the measured data and calculated values for erosion rate constant s1 using Eq. 5.9. 99

Page 11: RP 2002 Final Report - Georgia

ix

Figure Page No.

5.9. Comparison of the measured data and calculated values for erosion rate constant s2 using Eq. 5.10. 100

6.1. Comparisons of streamwise velocity profiles (location F1, F2, …, F6 left to right) at relative elevations of 0.6, 0.4, and 0.2 times the depth from top row downward. 104

6.2. Comparisons of turbulence kinetic energy. 106

6.3. Flint River Bridge Layout (a), and measurement cross-sections (b). 107

6.4. Comparisons of streamwise velocity profiles at locations from S1 to S6 (left to right)

and at water depths of 0.45, 0.3, and 0.1 times the depth (top to bottom row). 108

6.5. Snapshot of the streamlines at the Flint River bridge, including the footings in the simulation. 109

6.6. Geometry of the four overlapped grids around a single bent of the Ocmulgee River bridge piers . 110

6.7. Visualization of instantaneous flow field in the vicinity of the bridge piers. 110

6.8. Time-averaged streamwise velocity profiles at locations S1, S2, and S3 (left to right) and at water depths of 0.5, 0.3, and 0.1 times the depth from top to bottom row. 111

6.9 Distribution of scour depth at the equilibrium state. 113

6.10 Flow patterns near river bed. 113

6.11 Comparison of scour in prototype and laboratory cross sections at Cornelia. 117

6.12. Comparison of velocities between field and laboratory measurement at given distances from left side of the central pier bent. 118

6.13. Comparison of field and laboratory measurement of scour depths and Sheppard’s equations for Chattahoochee River, near Cornelia, GA. 119

6.14. Comparison of scour cross sections at Flint River bridge from laboratory model and Tropical Storm Alberto. 121

6.15. Comparison of field and laboratory measurement of scour depths and Sheppard’s equations Flint River at Bainbridge, GA. 123

Page 12: RP 2002 Final Report - Georgia

x

LIST OF TABLES

Table Page No.

2-1. Flood-frequency discharge data for Cornelia. 13

2-2. Flood-frequency discharge data for Macon. 19

2-3. Flood-frequency discharge data for Bainbridge. 25

3-1. Model studies conducted. 35

3-2. Raw experimental data for Chattahoochee River bridge near Cornelia, GA. 44

3-3. Dimensionless experimental data for Chattahoochee River bridge near Cornelia, GA. 44

3-4. Raw experimental data for Flint River bridge at Bainbridge, GA. 45

3-5. Dimensionless experimental data for Flint River bridge at Bainbridge, GA. 45

5-1. Location of samples and description of physiographic regions. 84

A-1. Linear, piecewise linear and exponential critical shear stress values for the erosion rate vs. applied shear stress models. 136

A-2. Linear, piecewise linear and exponential slope coefficients for the erosion rate vs. applied shear stress models. 137

A-3. Results of erosion tests and soil property tests on Murray County sample. 138

A-4. Results of erosion tests and soil property tests on Towns County sample. 139

A-5. Results of erosion tests and soil property tests on Habersham County sample. 140

A-6. Results of erosion tests and soil property tests on Haralson County sample. 141

A-7. Results of erosion tests and soil property tests on Bibb County sample. 142

A-8. Results of erosion tests and soil property tests on Wilkinson County sample. 143

A-9. Results of erosion tests and soil property tests on Effingham County sample. 144

A-10. Results of erosion tests and soil property tests on Decatur County sample. 145

Page 13: RP 2002 Final Report - Georgia

xi

Table Page No.

A-11. Results of erosion tests and soil property tests on Berrien County sample. 146

A-12. Results of erosion tests and soil property tests on McIntosh County sample. 147

Page 14: RP 2002 Final Report - Georgia

1

CHAPTER 1. INTRODUCTION

MOTIVATION

Bridge scour is a significant transportation problem because of the monetary damage and

possible loss of life that it can cause when it results in bridge foundation failure. The damage

caused by tropical storm Alberto in Georgia in 1994 is a case in point. Tropical storm Alberto

dumped as much as 28 in. of rainfall in parts of central and southwest Georgia from July 3-7,

1994, and caused numerous bridge failures and highway closings as a result of the 100-yr flood

stage being exceeded at many locations along the Flint and Ocmulgee Rivers. Approximately 18

ft of foundation scour occurred at the U.S. Highway 82 crossing of the Flint River near Albany,

Georgia (Stamey 1996). Total damage to the Georgia Department of Transportation highway

system was approximately 130 million dollars (Richardson and Davis 2001). Bridge failures can

also lead to loss of life such as in the 1987 failure of the I-90 bridge over Schoharie Creek near

Albany, New York; the US 51 bridge over the Hatchie River in Tennessee in 1989; and the I-5

bridges over Arroyo Pasajero in California in 1995 (Morris and Pagan-Ortiz 1999).

Prevention of bridge scour damages and possible loss of life hinges on having the

capability of predicting expected bridge scour. Unfortunately, such predictions remain a

challenging problem because of the complex interaction of the river flow with the obstruction

presented by the bridge foundation and with the erodible bed of the river. As the approach flow

encounters a bridge foundation it rolls, via the action of the foundation-induced adverse pressure

gradient, to form multiple large-scale vortices whose axes may be parallel (horseshoe vortices) or

perpendicular (tornado and/or whirlpool-type vortices) to the bed. These unsteady, large-scale

flow structures in conjunction with a broad range of turbulent scales control the sediment

Page 15: RP 2002 Final Report - Georgia

2

transport in the vicinity of the foundation and are responsible for the growth of scour holes

(Raudkivi, 1986; Dargahi, 1990, 1989; Melville, 1997). The complexity of such scour-inducing

flow phenomena has hampered satisfactory analysis and prediction procedures in the past. In

addition, bridge scour prediction has to be applicable to extreme hydrologic events and widely

varying river characteristics.

Previous attempts at solving the problem of scour prediction have been piecemeal at best.

Currently, scour prediction is based on formulas that have been developed from laboratory

experiments in flumes having unrealistic geometry. Furthermore, these formulas have been

applied using unproven scaling techniques to extrapolate from the model to the prototype. In

addition to the shortcomings of previous laboratory work, field verification of predicted scour

has been very limited even though notable efforts have been made using the latest in mobile

instrumentation techniques by the USGS (Landers and Mueller 1996). Long-term monitoring of

the scour process at particular bridge sites, including multiple spatial points of scour-depth

measurement with simultaneous real-time velocity monitoring, is practically nonexistent. A

further deficiency in current scour prediction methodology is that the terrific strides made in the

last decade or so in computational fluid dynamics (CFD), both in terms of computing power and

new computational methods, have yet to be applied in any comprehensive way to the problem of

bridge scour, despite the success enjoyed by CFD techniques in other areas of engineering.

Parola et al. (1997) have concluded that an integration of field studies, numerical model studies,

and laboratory studies is needed to make a significant advancement in bridge scour prediction

methodology and scour countermeasure design.

In addition to deficiencies in previous scour studies due to the lack of an integrated

approach, scour prediction in Georgia is further complicated by the wide degree of variability in

Page 16: RP 2002 Final Report - Georgia

3

physiographic regions, from the piedmont to the coastal plain, that have very different geologic

and topographic characteristics. The sediment erodibility and the type of flow field driving the

scouring process are highly regional because of these differences in geology and topography.

This research addresses the need for improving current bridge scour prediction

methodologies with particular reference to the State of Georgia. A comprehensive research

approach, consisting of 3D numerical modeling integrated with laboratory and field

measurements, is implemented to resolve multiple problems of scour prediction. The overall

purpose of the research is to develop a prediction methodology that will substantially reduce the

expense of foundations for new bridges and avoid the costly replacement of some existing

bridges due to anticipated scour that is overestimated by current prediction procedures. This

report summarizes the results of Phase 1 of the research and outlines the remaining tasks for

completion of a proposed comprehensive scour prediction methodology in the upcoming Phase 2

of the research.

BACKGROUND

Several factors make bridge scour prediction in Georgia particularly challenging. The state of

Georgia includes a number of unique physiographic regions from the piedmont to the coastal

plain that have very different geologic and topographic characteristics. For example, one of the

most important parameters of any scour prediction technique is the critical shear stress for

initiation of motion. While critical shear stresses are relatively well known for cohesionless

sands and gravels, the contribution of clays to form cohesive sediments results in difficulties

with respect to characterization of bed stability. Furthermore, sediment bed consolidation and

layering of sediments of different types both result in stratified sediment beds that have stability

properties that change with depth as scour occurs. The wide variety of sediments in Georgia

Page 17: RP 2002 Final Report - Georgia

4

make this problem of particular interest when formulating scour prediction equations, and failure

to be able to predict bed stability parameters can lead to significant errors in the prediction of

equilibrium scour depth.

In addition to sediment variability, the differences in topography throughout Georgia lead

to very different flow characteristics that can vary from local abutment scour caused by a bridge

on a wide, heavily vegetated floodplain in the coastal plain to contraction scour on relatively

narrow streams in the piedmont, or to contraction scour caused by tidal bridges subject to storm

surges on the coast. With this degree of variety in flow situations, it becomes difficult for a

single hydraulic model, especially if it is one-dimensional, to adequately forecast the relevant

velocities and/or shear stresses that are driving the scouring process.

Previous approaches to developing scour prediction equations have relied on laboratory

models, which in some instances have not properly modeled the field situation, especially in the

case of bridge abutments located on the floodplain in compound channel flow. However,

previous experimental work at Georgia Tech sponsored by Georgia DOT and FHWA in a large

flume of compound section has resulted in an improved methodology for predicting abutment

scour in compound channels (Sturm and Janjua 1994; Sturm and Sadiq 1996; Sturm 1998,

1999a,b). This methodology relies on estimates of a discharge contraction ratio rather than a

geometric contraction ratio in order to account for flow redistribution between the approach and

bridge contraction sections that depends on the lateral variability in velocity and depth in a

natural channel in overbank flow. Questions remain, however, on the "scaling issue" of applying

results from laboratory models to the field. Even laboratory models with realistic cross-sectional

shapes use sediment sizes and flow rates to reproduce a sediment mobility parameter, V/Vc ,

which represents the ratio of the approach velocity to the critical value for initiation of motion.

Page 18: RP 2002 Final Report - Georgia

5

Froude numbers are maintained in the subcritical range, but exact reproduction of the Froude

number is not usually attempted because some laboratory results indicate that the Froude number

has a smaller effect than the sediment mobility parameter. This modeling strategy has never

been fully validated except on an ad hoc basis with comparisons between scour predictions,

which are based on laboratory scour formulas, and measured field scour. In most cases, the

available field data displays too much variability to provide definitive comparisons, and so the

scaling issue remains unresolved. The advantage of laboratory models is that a wide range of

flow conditions can be investigated in a controlled experimental design to isolate the effects of

individual independent variables on the scour process.

During the last decade many investigations have been conducted to collect and analyze

field scour data at bridges during flood conditions. These studies have helped to set limits on

laboratory-derived equations, and have provided insight into the range of complex hydraulic and

sediment conditions that are encountered in the field. However, bridge scour investigations

focused on field studies alone are limited in that they provide data for only a very few conditions

(floods) at a given site, or they rely on measurements of remnant scour holes without adequate

knowledge of the hydraulic conditions causing the scour. Field data sets do not permit engineers

to evaluate the scour effects of a series of floods in which causative parameters, such as velocity

and sediment type, are altered.

In this study, field measurements of bridge scour are made with fixed instrumentation

that records velocity and bed depth continuously at specific points; and mobile instrumentation

that is deployed during high-flow events to record bathymetry and velocity field data through a

bridge reach. Fixed instrumentation provides continuous time series data when crews are not

deployed to collect data. This information is vital to understand overall processes of scouring and

Page 19: RP 2002 Final Report - Georgia

6

infilling at a bridge site. A typical limitation to fixed instrumentation data sets is the inability to

distinguish local scour from contraction scour. Because fixed instrumentation is always on a flow

obstruction, such as a pier, it measures the change in bed elevation due to both local and

contraction scour. In order to quantify local scour, channel geometry around the local obstruction

is needed. In order to quantify contraction scour, one must have some measure of average bed

elevation changes from upstream of the hydraulic influence of the bridge, through the bridge to

downstream of the hydraulic influence of the bridge. This can be accomplished for larger streams

from boats deployed during flood events. A further advantage of mobile deployed measurements

(accompanying fixed instrumentation) is that detailed 3-dimensional velocity field data sets are

collected through the study reach. These are particularly valuable because of the 3-dimensional

numerical modeling that is part of this investigation. Detailed field data sets, such as those

collected in this study, have been the most informative field data collected for understanding

scour processes (Landers and Mueller 1996).

Both field and laboratory data collection have limits in the range of scenarios that can be

observed. This problem can be overcome by use of a numerical model of bridge scour, having

adequate complexity and computational accuracy to represent complex scour processes; and

having adequate observed laboratory and field data to permit calibration and refinement to

represent reality. A three-dimensional numerical model using modern Computational Fluid

Dynamics (CFD) techniques is needed to fully understand the complexity of bridge foundation

flows. CFD methods have been successfully applied to a number of real-life engineering

problems (Neary et al. 1999; Sinha et al. 1998) but have yet to be utilized to their full potential to

tackle the scour problem. A few preliminary simulations have been reported in the literature

(Olsen and Melaan 1993; Dou et al. 1996), but they all lack the level of modeling rigor and

Page 20: RP 2002 Final Report - Georgia

7

sophistication required to reproduce the multiple facets of the phenomenon. More specifically,

these studies have adopted very simplistic turbulence modeling strategies and have, thus, failed

to reproduce even the broad qualitative features of the flow in the vicinity of the foundation.

Recently, however, very sophisticated turbulence models, capable of quantitatively accurate flow

predictions, have been developed and successfully applied to calculate a wide range of complex,

3D flows with vortices (Sotiropoulos and Patel 1994, 1995a,b; Sotiropoulos and Ventikos 1998).

The first attempt to apply such a model to bridge abutment flows was reported by Sotiropoulos et

al. (1999). Their numerical results provided the first in-depth insights into the complexities of

bridge foundation flows and demonstrated the feasibility of advanced CFD modeling for

developing predictive models of scour.

It is important to point out that fully three-dimensional CFD models, such as the one

developed by Sotiropoulos et al. (1999), promising as they may be, are presently suitable only

for academic computations. This is because the computational times they require are, even with

today’s very powerful supercomputers, far too excessive for them to be useful as engineering

design tools. Yet such models provide the only alternative for obtaining the in-depth

understanding of the physics of bridge-foundation flows needed for refining the predictive

capabilities of simpler engineering models—such as the 1D models used extensively in the

design of bridge foundations today. It is in this spirit that an advanced three-dimensional CFD

model, employed in conjunction with field and laboratory studies, can be an invaluable

engineering tool that can make an immediate and very significant contribution to the present-day

state of the art.

Page 21: RP 2002 Final Report - Georgia

8

RESEARCH OBJECTIVES

The overall objective of the research described in this report is to improve bridge scour

predictions using one-dimensional (1D) methods by combining physical modeling in the

laboratory, field monitoring, and three-dimensional (3D) numerical modeling. The specific

objectives of the research are:

(1) Monitor selected bridges in the field using fixed instrumentation, mobile

deployments of instrumentation from boats, and historic scour-surface evaluation;

(2) Physically model in the laboratory two particular bridge sites that are also

undergoing detailed field monitoring, validate the physical models, and test the

models over a wide range of flow conditions;

(3) Apply a 3D numerical model, which is being developed in separate research

sponsored by NSF, to two existing bridge sites having detailed instrumentation for

validation of the model with field and laboratory data collected in this research;

(4) Apply the 3D numerical model to simulate the flow fields at additional bridge sites

that have more limited instrumentation;

(5) Characterize sediments in general physiographic regions of Georgia in terms of their

erodibility;

(6) Combine the knowledge gained from the field, laboratory, and numerical modeling

into an improved 1D scour prediction methodology specific to various regions of

Georgia.

Page 22: RP 2002 Final Report - Georgia

9

ORGANIZATION OF THE REPORT

The results of the field scour data collection program, the laboratory model studies, and the 3D

numerical model studies are given in Chapters 2, 3, and 4, respectively. Chapter 5 summarizes

the results of flume erodibility tests made on Shelby tube sediment samples collected by GDOT

at 10 representative bridge sites in Georgia. Comparisons of laboratory and field results for

velocity and scour depths as well as detailed comparisons of velocity and turbulence fields

measured in the laboratory with those predicted by the numerical model are discussed in detail in

Chapter 6. Finally, a summary and conclusions are provided in Chapter 7 in which the unique

observations associated with combined field, laboratory, and numerical model studies are

summarized, and their implications with respect to a proposed scour prediction methodology to

be completed in Phase 2 of the research are discussed.

Page 23: RP 2002 Final Report - Georgia

10

CHAPTER 2. FIELD DATA COLLECTION

FIXED FIELD INSTRUMENTATION

Fixed field instrumentation has been installed at four bridge sites, which represent

various sediment types and physiographic regions in Georgia. USGS gaging stations are located

at these sites. Detailed fixed instrumentation has been installed at two of the sites. One of the

detailed field instrumentation sites is located in the coastal plain (Bainbridge), and the second

site is located in the Piedmont Province (Macon). Less detailed fixed instrumentation has been

installed at the remaining two sites (Darien and Cornelia). The detailed sites have the following

equipment:

• stage sensor;

• cross-channel two-dimensional velocity sensor;

• fathometer array to record streambed elevation;

• raingage;

• data logger and controller for each device;

• solar panel and instrumentation shelter; and

• satellite telemetry.

The less detailed sites have the same equipment except for the velocity sensor.

The fathometers are attached to the bridge piers in order to monitor the changes in bed

elevation around the bridge pier. Water velocity is also a critical bridge scour parameter that is

used to quantify the available scour energy. The cross-channel velocity sensor provides two-

dimensional velocity for a series of points across the channel in the bridge-approach section. The

sensor is mounted at a fixed location and aimed across the channel. The velocity meter uses

Page 24: RP 2002 Final Report - Georgia

11

acoustic-Doppler technology and has its own system controller on site. Velocities are recorded at

15-minute intervals.

MOBILE FIELD INSTRUMENTATION

A mobile scour data-collection system has four components: instruments to measure

velocity and channel-geometry data; instruments to deploy equipment in the water; an instrument

to measure the horizontal position of the data collected; and a data storage device. For this

investigation, an acoustic Doppler current profiler (ADCP) is deployed from a manned boat and

used to measure three-dimensional velocity profiles. A recording digital fathometer is used to

measure channel depths. Horizontal position is measured using a kinematic differential Global

Positioning System (GPS). Some of the parameters collected with the mobile instrumentation

include:

• detailed channel geometry at and near the bridge;

• approach-flow velocities over the study reach;

• water-surface slope during flood events;

• visual analysis and notes on the surface velocity direction, channel and overbank

roughness, and vegetation cover;

• approximate measurements of the extent and composition of debris;

• photographs of channel and bridge at flood and low-flow conditions;

• water temperature;

• bridge and pier geometry; and

• bed sediment samples and soil boring logs from the bridge crossing.

All data is recorded and used to interpret and extend the data collected by the fixed

instrumentation, and for the 3D numerical modeling component of this investigation.

Page 25: RP 2002 Final Report - Georgia

12

STREAMBED SEDIMENT SAMPLING

Bed-material characteristics are important determinants of streambed erodibility and bed-

material transport conditions. The objective for any of the collection techniques is to ensure that

a representative sample is collected. The BMH-53 or BMH-80 hand samplers are used to collect

the samples in sand-bed streams that can be waded. A BM-54 is used to collect samples in sand-

bed streams that are too deep to be waded. Procedures are not well defined for sampling

cohesive bed-materials, but a BMH-53 or similar cylinder sampler may be used on streams that

can be waded. The type of sampler used will always be noted with the bed-material data.

Sampling locations were selected to ensure samples are representative of the bed material

controlling the sediment-transport processes in the study reach. In streams with cohesive beds,

sediment in the zone of scour was sampled. Bed material samples were collected from several

locations both in the bridge approach and the bridge sections, including in local scour holes.

Bed-material samples were analyzed by the Georgia Institute of Technology laboratories for

grain-size distribution and other properties related to bridge scour.

FOUR SELECTED SITES

Chattahoochee River Near Cornelia, Ga (02331600)

Site Description

The first site chosen for the research project is the Chattahoochee River near Cornelia,

Georgia. The USGS has been gaging stage and streamflow at this site since 1957. The gage is

located at the Georgia Highway 384 (Duncan Bridge Road) bridge at the Habersham-White

County line in Northeast Georgia. The drainage area at this site is 315 square miles. The

channel in the vicinity of the bridge is a long quiet pool about 160 feet wide at lower flows. The

pool extends from the rock ledge control about 500 feet downstream of the bridge to about 3,000

Page 26: RP 2002 Final Report - Georgia

13

feet upstream of the bridge. The banks will overflow at higher stages onto a flat but narrow

floodplain. The control is a rock ledge, which runs diagonally from the right bank downstream to

the left bank. The bridge piers consist of four concrete square columns, which rest on concrete

footings buried below the streambed (Fig. 2.1). There is one bridge pier located in the center of

Figure 2.1. Chattahoochee River near Cornelia, GA.

the channel and one bridge pier on each of the banks. The bridge piers are aligned with the flow.

The peak discharge of record is 26,400 cubic feet per second, which occurred on March

12, 1963. The 2-year flood event at this site is about 11,800 cubic feet per second, and the 500-

year flood event is about 39,600 cubic feet per second as shown in Table 2-1.

Table 2-1. Flood-frequency discharge data for Cornelia.

Recurrence Interval (years)

Discharge (cfs)

2 11,800 10 21,500 50 29,400

100 32,600 500 39,600

Page 27: RP 2002 Final Report - Georgia

14

Fixed-Field Instrumentation

The fixed-field instrumentation at this site consists of four fathometers, an acoustic

velocity meter, a raingage, and a stage sensor. The acoustic velocity meter, raingage, and stage

sensor are interfaced with a Data Collection Platform (DCP), which logs readings from the

sensors every 15 minutes. The DCP transmits the 15-minute data from each of the sensors every

four hours using satellite telemetry. The four fathometers are interfaced with a data logger, which

logs the readings from the fathometers every 30 minutes. The fathometers are attached to the

center bridge pier and monitor the bed elevation changes occurring around the bridge pier. The

velocity meter is attached to the nose of the center bridge pier and monitors two-dimensional

velocities at three points across the approach bridge section (Fig. 2.2).

Figure 2.2. Fathometer and velocity meter layout at Cornelia.

Data Collected

Due to drought conditions during the 2-year study, data was collected during five

moderate highwater events at the Cornelia site. The peak discharges for these events ranged from

3,400 to 13,600 cfs. The 13,600 cfs event is slightly greater than the 2-year recurrence interval

for this site. The event resulted in two additional feet of scour in comparison to the pre-existing

Page 28: RP 2002 Final Report - Georgia

15

scour hole at the nose of the pier as shown in Fig. 2.3. The peak velocity recorded by acoustic

velocity meter during the July 2003 event was approximately 7 feet per second. The event of

July 2003 peaked early in the morning. The stage was already falling when the crew with the

mobile instrumentation arrived, and the boat and ADCP could not be deployed in the water due

to the falling stage. However, velocity distribution and cross-section profile data were collected

from the upstream side of the bridge as the stage was falling.

Figure 2.3. Fathometer data from Cornelia.

Chattahooche River near Cornelia, GAJuly 1-3, 2003

1120

1121

1122

1123

1124

1125

1126

6 6.5 7 7.5 8 8.5 9

Time (days)

Bed

Elev

atio

n (fe

et)

0

2000

4000

6000

8000

10000

12000

14000

Disc

harg

e (c

fs)

Page 29: RP 2002 Final Report - Georgia

16

The 13,600 cfs event that occurred in July 2003 was modeled in the hydraulics laboratory

at Georgia Tech. The results from the laboratory models are shown in Chapter 3, and the data

comparisons between the field and model data are shown in Chapter 6. A historic discharge

measurement was made at the site in 1961 at a discharge of 13,100 cfs, which is very close to the

peak discharge of the July 2003 event. A cross-section comparison was made at the upstream

side of the bridge for both the July 2003 and December 1961 event as shown in Fig. 2.4. The

maximum scour that occurred during the July 2003 event was within one foot of the scour that

occurred during the discharge measurement that was made in December of 1961. The shapes of

the scour hole around the center bridge pier were also similar for both events indicating

recurrence of the same extent of scour for the same discharge after repeated infilling between

floods.

1110

1115

1120

1125

1130

1135

50 100 150 200 250 300

Width (feet)

Bed

Ele

vatio

n (f

eet)

12-Dec-61 13-Jun-03 2-Jul-03

Figure 2.4. Cross-section comparison at the upstream side of the bridge at Cornelia.

Page 30: RP 2002 Final Report - Georgia

17

Eleven cross-sections of the river and floodplain were surveyed along the channel reach

near the bridge as illustrated in Fig. 2.5. The position of the bridge piers was determined using a

differential Global Positioning System (GPS). The cross-sections and pier placement data were

used to construct the laboratory and three-dimensional computer models. Bed material samples

were also collected at three of the cross-sections. The particle size distributions of the collected

bed material samples are shown in Fig. 2.6.

Figure 2.5. Layout of surveyed cross-sections at Cornelia.

Ocmulgee River At Macon (02213000)

Site Description

The second site chosen for the project is the Ocmulgee River at Macon, Georgia. The

USGS has been gaging stage and streamflow at this site since 1895. The gage is located at the

Fifth Street Bridge (Otis Redding Bridge) in Macon. The drainage area at this site is 2,240

square miles. The channel upstream of the bridge is straight for about 1,000 feet and straight for

Page 31: RP 2002 Final Report - Georgia

18

Figure 2.6. Particle size distribution of bed material samples at Cornelia.

about 1,500 feet downstream of the bridge. The streambed is smooth and sandy. The right bank

is high and is not subject to overflow. The left bank is subject to overflow at high stages, but the

highway fill confines all flow to the bridge opening. The control is a shifting sand streambed.

The bridge piers consist of four cylindrical columns that rest on concrete footings, which are

buried below the streambed. As shown in Fig. 2.7, there is one bridge pier in the center of the

channel one bridge pier at each of the banks. All three bridge piers are aligned with the flow.

The peak discharge of record is 107, 000 cubic feet per second, which occurred on July 6,

1994. The 2-year flood event is about 28,500 cubic feet per second, and the 500-year flood event

is about 108,400 cubic feet per second as shown in Table 2-2.

PARTICLE-SIZE DISTRIBUTION (Chattahoochee River)

0

10

20

30

40

50

60

70

80

90

100

0.010.1110

Particle Size (mm)

Perc

ent P

assi

ng (%

)

Section 10-3/3 Section 10-Right Bank Section 10-Left Bank Section 7-3/3Section 7-2/3 Section 7-1/3 Section 2-2/3 Section 2-1/3

Page 32: RP 2002 Final Report - Georgia

19

Figure 2.7. Ocmulgee River at Macon, GA.

Table 2-2. Flood-frequency discharge data for Macon.

Recurrence Interval (years)

Discharge (cfs)

2 28,500 10 56,200 50 79,200

100 88,300 500 108,400

Page 33: RP 2002 Final Report - Georgia

20

Fixed-Field Instrumentation

The fixed-field instrumentation at the Macon site consists of six fathometers and a stage

sensor. The stage sensor is interfaced with a DCP, which transmits 15-minute data from each of

the sensors every hour. The six fathometers are interfaced with a data logger, which logs the

readings from the fathometers every 30 minutes. Five fathometers are attached to the center

bridge pier, and one fathometer is located at the nose of the pier on the right bank (Fig. 2.8).

Figure 2.8. Fathometer layout for Ocmulgee River at Macon, GA.

Data Collected

Data was collected during multiple moderate highwater events. The highest peak

discharge of these events was 25,500 cfs, which is below the 2-year occurrence interval for this

site. During this 25,500 cfs event in May of 2003, three additional feet of scour occurred around

the center bridge pier as shown in Fig. 2.9. During a smaller event in March 2003, velocity

distribution data were collected at various cross-sections throughout the channel reach using an

Acoustic Doppler Current Profiler (ADCP). The peak discharge during this March 2003 event

was 20,300 cfs.

Page 34: RP 2002 Final Report - Georgia

21

Figure 2.9. Fathometer data from Macon.

A historic discharge measurement was made at the upstream side of the bridge during a

highwater event in March of 1998. The peak discharge during this event was 65,000 cfs. The

cross-section from this historic measurement was compared with cross-sections collected at the

upstream side of the bridge during this study as shown in Fig. 2.10. The historic cross-section

Ocmulgee River at Macon, GAMay 1-18, 2003

265

266

267

268

269

270

0 2 4 6 8 10 12 14 16 18Time (days)

Bed

Ele

vatio

n (f

eet)

500

5500

10500

15500

20500

25500

Dis

char

ge (c

fs)

Page 35: RP 2002 Final Report - Georgia

22

shows 10 feet of contraction scour. The peak velocity measured during this event was nearly 10

feet per second.

Figure 2.10. Cross-section comparison at the upstream side of the bridge at Macon.

Seven cross-sections were surveyed throughout the channel reach and were used to

construct the three-dimensional model (Fig. 2.11). Bed material samples were taken at four of

the cross-sections. The particle size distributions of the collected bed material samples are shown

in Fig. 2.12.

240

250

260

270

280

290

300

310

320

330

-20 30 80 130 180 230 280 330 380 430

Station (feet)

Bed

Ele

vatio

n (f

eet)

22-Mar-03 21-May-03 21-Jul-03 10-Mar-98

Page 36: RP 2002 Final Report - Georgia

23

Figure 2.11. Layout of surveyed cross-sections at Macon.

Figure 2.12. Particle size distribution of bed material samples at Macon.

OCMULGEE RIVER AT MACON, GEORGIACUMULATIVE PARTICLE-SIZE PLOT

0%

10%

20%

30%

40%

50%

60%

70%

80%

90%

100%

0.010.1110Particle Size (mm)

Perc

ent P

assing

Cross Section #1 Cross Section #2Cross Section #3 Cross Section #4

Page 37: RP 2002 Final Report - Georgia

24

Flint River At Bainbridge (02356000)

Site Description

The third site chosen for the research project is the Flint River at Bainbridge, Georgia.

The USGS has been gaging stage and streamflow at this site since 1908. The gage is located at

the Business Highway Route 27 Bridge in Bainbridge. The drainage area at this site is 7,570

square miles. The channel is fairly straight for several thousand feet upstream and has a sharp

bend about 500 feet downstream. The site is affected by backwater from the Jim Woodruff

Reservoir at lower stages. At higher stages, the backwater is negligible, and the banks overflow

onto a very wide and flat floodplain. The bridge piers consist of two square concrete columns

that rest on very large square concrete footings (Fig. 2.13). The large footings protrude from the

streambed. There are two bridge piers in the main channel and both are aligned with the flow.

Figure 2.13. Flint River at Bainbridge, GA.

The peak discharge of record is 108,000 cubic feet per second, which occurred on July

14, 1994. The 2-year flood event is about 30,900 cubic feet per second, and the 500-year flood

event is about 122,000 cubic feet per second as given in Table 2-3.

Page 38: RP 2002 Final Report - Georgia

25

Table 2-3. Flood-frequency discharge data for Bainbridge.

Recurrence Interval (years)

Discharge (cfs)

2 30,900 10 58,100 50 83,700

100 94,900 500 122,000

Fixed-Field Instrumentation

The fixed-field instrumentation at this site consists of seven fathometers, a raingage, two

acoustic velocity meters, and a stage sensor. The stage sensor, raingage, and acoustic velocity

meters are interfaced with a DCP, which logs readings from the sensors every 15 minutes. The

DCP transmits the 15-minute data from each of the sensors every four hours using satellite

telemetry. The seven fathometers are interfaced with a data logger, which logs the readings from

the fathometers every 30 minutes. Four fathometers are attached to the left center bridge pier,

and three fathometers are attached to the right center bridge pier (Fig. 2.14). The fathometers

monitor the change in bed elevation around the bridge piers. An acoustic velocity meter is

attached to the nose of the left center pier, and measures two-dimensional velocities at three

points across the approach bridge section. A second acoustic velocity meter is attached to the

downstream end of the left center bridge pier. This velocity meter measures a one-dimensional

velocity at point in the center of the main channel. This one-dimensional velocity is used as an

index velocity to determine the discharge.

Page 39: RP 2002 Final Report - Georgia

26

Figure 2.14. Fathometer and velocity meter layout at Bainbridge.

Data Collected

Due to the drought conditions during the study, no highwater events occurred at this site.

So there were no pertinent data collected at this site. However, a historic discharge measurement

was made during the peak discharge of record in 1994. The discharge measurement contained a

velocity distribution and cross-section elevations at the upstream side of the bridge. Georgia

Tech physically modeled this event, and the velocity distribution and cross-section compared

very well as shown in Chapter 6. Several historic cross-sections are shown in Fig. 2.15, including

the cross-section obtained during the peak discharge of record in July of 1994. The cross-section

from the 1994 flood shows as much as 10 feet of scour when compared with previous and recent

cross-sections to identify the bed elevation that would occur in the absence of local scour.

Five cross-sections were surveyed and used to construct the model. The railroad bridge

upstream of the bridge was also surveyed and used in the modeling (Fig. 2.16). Bed material

samples were collected at four cross-sections, and the particle size distribution of the collected

bed material samples are shown in Fig. 2.17.

Page 40: RP 2002 Final Report - Georgia

27

40

50

60

70

80

90

100

0 100 200 300 400 500 600 700 800 900

Width, feet

Elev

atio

n, fe

et

10-Mar-80 14-Jul-94 12-Jul-94 21-Mar-01 20-Mar-02

Figure 2.15. Cross-section comparison at Bainbridge, GA.

Figure 2.16. Layout of surveyed cross-sections at Bainbridge.

Page 41: RP 2002 Final Report - Georgia

28

Figure 2.17. Particle size distribution of bed material samples at Bainbridge at left (L), right (R), and center (C) of main channel for cross section numbers shown in Fig. 2.16.

Darien River At Darien (02203598)

Site Description

The fourth site chosen for the project is the Darien River at Darien, Georgia. The USGS

began gaging stage and streamflow at this site in January 2002. The gage is located at the

Georgia Highway 17 bridge in Darien. The channel is fairly straight in the vicinity of the bridge.

The flow at this site is tidally affected and is confined in a 400 foot wide channel. The bridge

piers consist of three rectangular concrete columns, which rest on very large square concrete

footings (Fig. 2.18). The footings protrude from the streambed. The mid-sections of the columns

PARTICLE-SIZE DISTRIBUTION (Flint River)

0%

10%

20%

30%

40%

50%

60%

70%

80%

90%

100%

0.0010.010.1110

Particle Size (mm)

Perc

ent P

assin

g

2R - No material 2C 2L 3R3C 3L 4R 4C4L 5R 5C 5L

Page 42: RP 2002 Final Report - Georgia

29

are connected with a flange. There are four bridge piers in the main channel. Two of the bridge

piers near the right bank are protected with wooden bridge fenders while the others are not.

Figure 2.18. Darien River at Darien, GA.

Fixed-Field Instrumentation

The fixed-field instrumentation at this site consists of seven fathometers, a raingage, and

a stage sensor. The stage sensor and raingage are interfaced with a DCP, which logs readings

from the sensors every 15 minutes. The DCP transmits the 15-minute data from each of the

sensors every four hours using satellite telemetry. The seven fathometers are interfaced with a

data logger, which logs the readings from the fathometers every 30 minutes. Three fathometers

are attached to the left bridge fender, two fathomers are attached to the right bridge pier, and two

fathometers are attached to the left center pier (Fig. 2.19). The fathometers monitor the change in

bed elevation around the bridge fenders and piers.

Page 43: RP 2002 Final Report - Georgia

30

Figure 2.19. Fathometer layout at Darien.

Data Collected

Fathometer data were continuously collected at Darien. During the tide cycle, one foot of

scour and fill was seen on a few occasions at a couple of the fathometer locations as shown in

Fig. 2.20. The scour and fill coincided with the tide cycle. Overall, the scour and fill seen on a

daily basis was minimal. However, on a yearly basis the scour and fill was as much as five feet at

one of the fathometer locations (Fig. 2.21).

Nine cross-sections throughout the channel reach were surveyed and used to construct the

models (Fig. 2.22). The position and dimensions of the bridge fenders were also surveyed and

incorporated into the models. Bed material samples were taken at five locations near the bridge,

and the particle size distribution of these bed material samples are shown in Fig. 2.23.

Page 44: RP 2002 Final Report - Georgia

31

Figure 2.20. Fathometer data from Darien.

Figure 2.21. Fathometer data at downstream side of left bridge fender.

Darien River at Darien, GAMay 29, 2002 - June 2, 2002

-19.0

-17.0

-15.0

-13.0

-11.0

-9.0

-7.0

-5.0

-3.0

-1.0

130 130.5 131 131.5 132 132.5 133 133.5 134 134.5 135Time (days)

Bed

Ele

vatio

n (fee

t)

4

6

8

10

12

14

Gag

e Hei

ght (

feet

)

Page 45: RP 2002 Final Report - Georgia

32

Figure 2.22. Layout of surveyed cross-sections at Darien.

Figure 2.23. Particle size distribution of bed material samples at Darien.

PARTICLE-SIZE DISTRIBUTION (Darien River)

0%

10%

20%

30%

40%

50%

60%

70%

80%

90%

100%

0.010.1110 Particle Size (mm)

Perc

ent P

assi

ng

Downstream Fender Upstream Fender Downstream Bridge Upstream Bridge Front of Pier

Page 46: RP 2002 Final Report - Georgia

33

CHAPTER 3. LABORATORY MODEL STUDIES

INTRODUCTION

Although numerous formulas for the prediction of bridge pier scour depths have been developed

based on laboratory experiments as summarized by Melville and Coleman (2000) or Sturm

(2001), for example, considerable doubt remains concerning their applicability to large-scale

prototypes. Scour-depth estimates based on laboratory data tend to overestimate actual pier scour

depths measured in the field (Landers and Mueller 1996). This situation is partly due to the

sediment scale effect that limits the size of the sediment that can be used in the laboratory

without it becoming so fine-grained that interparticle forces that may not exist in the field

become dominant in the laboratory (Ettema et al. 1998).

This chapter discusses laboratory modeling considerations with respect to bridge scour

and gives the results of laboratory scale model studies conducted in the Georgia Tech Hydraulics

Laboratory. The two field sites that were modeled in the laboratory are the Chattahoochee River

bridge near Cornelia, Georgia and the Flint River bridge at Bainbridge, Georgia which were

described in Chapter 2. Each site was modeled at two different physical scales for two sediment

sizes. Both bank-full and extreme flood flows were modeled.

MODELING CONSIDERATIONS

Dimensional analysis of the pier scour problem produces (Ettema et al. 1998, Sturm 2001):

),,,,,( 1

50

1 FrVV

db

by

KKfbd

cs

sθ= (3.1)

in which ds = scour depth; b = pier width; Ks = shape factor; Kθ = skewness factor; y1 = approach

depth; V1 = approach velocity; Vc = critical velocity; d50 =median sediment size; and Fr =

approach Froude number. For strict dynamic similarity, all of the dimensionless parameters on

Page 47: RP 2002 Final Report - Georgia

34

the right-hand side of Eq. 3.1 should be the same in model and prototype. However, as a

practical matter, choosing a length scale ratio such that the model will fit in a laboratory results

in the impossible situation of d50 becoming so small that a cohesive sediment would be necessary

to satisfy equality of the ratio b/d50 . A similar dilemma arises in the case of modeling

contraction scour. To further complicate matters, a sand-bed river is likely to be in the live-bed

scour regime (V1/Vc > 1) which is difficult to model in the laboratory because of the sediment

that must be fed or recirculated to the flume in order to maintain the unknown upstream sediment

transport rate.

The modeling compromise utilized in this study was to maintain approximate Froude

number similarity with equality of y1/b values while selecting the sediment size such that clear-

water scour was obtained near the maximum of V1/Vc = 1.0 as a conservative approach. The

experimental plan proceeded in two stages with two different physical model scales for each

bridge. In the initial stage, a flat mobile-bed model was constructed at a larger scale for the

purpose of making detailed velocity and turbulence measurements around the main pier bent for

comparison with the numerical model results and for measuring scour depths to be compared

with accepted scour prediction formulas. These initial-stage models, referred to as flat-bed

models, were intended to reproduce bank-full flows rather than flood flows, and the Froude

number was allowed to take on larger values than occurred in the prototype. In the second stage

of the model studies, smaller-scale models referred to as river models were constructed to

include the complete river bathymetry and bridge geometry for reproducing both bank-full and

extreme flood events (100-yr and 200 yr). A smaller sediment size was used to achieve close

Froude number similarity in the range of V1/Vc = 0.7 – 1.0. A summary of the model study

conditions is given in Table 3-1.

Page 48: RP 2002 Final Report - Georgia

35

Table 3-1. Model studies conducted.

Bridge Model Type of Model Scale of Model Sediment Size, mm Flow Conditions Chattahoochee R. near Cornelia

Flat Bed 1:23.3 3.3 mm Bank-full

Chattahoochee R. near Cornelia

River Model 1:40 3.3 mm, 1.1 mm Bank-full, 100-yr

Flint R. at Bainbridge

Flat Bed 1:50 3.3 mm Bank-full

Flint R. at Bainbridge

River Model 1:90 1.1 mm Alberto (200-yr)

EXPERIMENTAL METHODS

Model Construction

Experiments were conducted in a 4.2 m wide by 24.4 m long flume with a mobile sediment bed

in which models of the Chattahoochee River bridge near Cornelia, Georgia and of the Flint River

bridge at Bainbridge, Georgia were constructed. Two median sediment sizes of d50 = 3.3 mm and

1.1 mm were utilized. Each sediment was relatively uniform in size distribution with a geometric

standard deviation, σg = 1.3.

In the flat-bed model studies, the mobile bed was leveled between temporary walls of

cinder blocks to form a flat bed with a channel width of 8.0 ft and a working channel length of

60 ft in which only the main-channel pier bent was placed. In the river model studies, the

complete river bathymetry was modeled with a fixed-bed approach channel followed by a

mobile-bed working section in which the bridge embankment and pier models were placed. In

the river models, the approach section was approximately 40 ft long with a 5-ft long working

section and then an approximately 10-ft long exit section for sediment deposition. The river

bathymetry was modeled by cutting plywood templates that reproduced the surveyed cross-

sections that were shown in Chapter 2 and then leveling the bed to match the templates. The

Page 49: RP 2002 Final Report - Georgia

36

templates were left in place in the fixed-bed section, but in the moveable-bed section they were

installed and removed after the bed was shaped for each experimental run. Fiberglass was used

for the fixed bed in the Chattahoochee River model and a single layer of the 3.3-mm gravel was

fixed to it by spraying it with several layers of polyurethane. For the Flint River model, the full

depth of the 3.3-mm gravel bed was simply molded to the plywood templates and a surface layer

was fixed with polyurethane. This proved to be a more flexible scheme for moving different

models in and out of the flume.

The initial velocity and turbulence measurements were made for a fixed bed in the

vicinity of the model pier bent. This was achieved by spraying polyurethane on the gravel to hold

it in place. For subsequent scour experiments, the fixed-bed section was removed and the bed

was made completely mobile, but no scour occurred upstream of the pier bent because conditions

for incipient live-bed scour were not exceeded.

The central pier bent for the Cornelia bridge is shown in Fig. 3.1 with prototype

dimensions. It was modeled at a scale of 1:23.3 for the flat-bed model and at a scale of 1:40 for

the river model as shown previously in Table 3-1. The inner piers are tapered as shown in the

figure. They are the original piers in existence before widening of the bridge occurred. The outer

rectangular piers, which have a width of 3.5 ft and a length of 4.0 ft in the flow direction, were

added when the bridge was widened. The spacing between the two inner piers is 16.0 ft, and

between the outer and inner piers it is 6.75 ft. The inner piers are connected by a solid web that

extends from an elevation above low-water stage to an elevation near the 100-yr high-water

stage. The footings were also modeled at the same scale as the piers and placed at the correct

elevation relative to the channel bed as shown in Fig. 3.1.

Page 50: RP 2002 Final Report - Georgia

37

Upstream

55.8 ft

44.5 ft

17.0 ft

Top of the bed (EL= 1126 ft)

2yr flood flow(Q=13500 cfs)

100 yr flood flow(Q=31700 cfs)

24.8 ft

14.0 ft5.0 ft

6.75 ft 16.0 ft 11.75 ft

10.0 ft11.5 ft3.5 ft 2.0 ft2.0 ft 14.0 ft

C

C'

D

D'

B

B'

A

A'

(a) Profile.

SEC.A-A' SEC.B-B' SEC.C-C' SEC.D-D'

13.75 ft

3.5 ft

2.25 ft

53.55 ft

10.67 ft

3.5 ft

48.85 ft

2.75 ft

3.17 ft 3.17 ft

2.25 ft

13.75 ft

50.44 ft

3.5 ft

2.0 ft

46.95 ft

2.75 ft

10.67 ft

Top of the bed (EL= 1126 ft)

14.0 ft

14.0 ft

24.8 ft

100 yr flood flow(Q=31700 cfs)

2yr flood flow(Q=13500 cfs)

(b) Sections.

Figure 3.1. Sketch of central pier bent of Chattahoochee River bridge near Cornelia, GA with prototype elevations and dimensions.

Page 51: RP 2002 Final Report - Georgia

38

The main channel pier bent of the Flint River bridge at Bainbridge is shown in Fig. 3.2. It

was modeled at a scale of 1:50 for the flat-bed model and at a scale of 1:90 for the river model.

The pier bent consists of two identical piers that are 6 ft square in cross section with a spacing in

the flow direction of 49 ft centerline to centerline. The footings are stepped as shown in the

figure. The high water flood that was modeled for this case occurred due to Tropical Storm

Alberto in 1994 with an estimated return interval of 200 years.

Photographs of the river models for the Chattahoochee River bridge at a 1:40 scale and

the Flint River bridge at a scale of 1:90 are shown in Figs. 3.3 and 3.4, respectively. The riprap

on the embankments is shown in Fig. 3.3 for the Chattahoochee River bridge. The central pier

bent in the middle of the river is the only one that experienced any significant scour, and so it is

the bent for which scour results are given subsequently. The left floodplain can be seen just

downstream of the bridge and it is relatively narrow there as well as upstream of the bridge. The

downstream rock sill that serves as a control for low flows as discussed in Chapter 2 was also

modeled although it is not shown in the photograph.

The four bents of the Flint River bridge can be seen in Fig. 3.4 (a) as well as the old

embankment on the left side of the bridge opening which remained in place after the new bridge

was built. This embankment blocked the flow around the leftmost pier but directed some flow

toward the pier that is second from the left in the photograph. However, the third pier from the

left is located in the deepest part of the main channel, and it experienced the greatest scour

depths which are reported subsequently in this chapter. The upstream railroad bridge was

modeled and is shown in Fig. 3.4 (b). The left embankment of this bridge effectively blocked

much of the approach flow to the highway bridge in the left floodplain. Flow visualization

showed that flow in the right floodplain did not affect scour around the primary pier in the study.

Page 52: RP 2002 Final Report - Georgia

39

Upstream

Top of the bed (EL= 57.6 ft)

Bankfull flow(Q=45000 cfs)

Alberto (7/12/94)(Q=107000 cfs)

43.1 ft

53.3 ft76.3 ft

55.0 ft

6.0 ft

15.7 ft

A

A'

(a) Profile.

Top of the bed (EL= 57.6 ft)

Bankfull flow(Q=45000 cfs)

Alberto (7/12/94)(Q=107000 cfs)

43.1 ft

53.3 ft

15.7 ft10.0 ft

21.8 ft

18.8 ft

6.0 ft

SEC. A-A'

(b) Section.

Figure 3.2. Sketch of central pier bent of Flint River bridge at Bainbridge with prototype elevations and dimensions.

Page 53: RP 2002 Final Report - Georgia

40

(a) Bridge section looking downstream (b) River and bridge looking downstream from the right bank. from left floodplain.

Figure 3.3. Model of Chattahoochee River bridge near Cornelia.

(a) Bridge section looking downstream from (b) Railroad bridge looking downstream left bank including old embankment. to highway bridge.

Figure 3.4. Model of Flint River bridge at Bainbridge.

Page 54: RP 2002 Final Report - Georgia

41

Experimental Instrumentation

The water supply to the flume was provided from a large constant-head tank through a 12-in.

diameter pipe that can deliver up to 10 ft3/s to the head box of the flume. A flow diffuser,

overflow weir, and baffles in the flume head box provided stilling of the inflow to reduce

entrance effects and produce a uniform flume inlet velocity distribution. A flap tailgate

controlled the tailwater elevation. Water recirculated through the laboratory sump from which

two pumps continuously provided overflow to the constant-head tank. In the 12-in. supply pipe,

discharge was measured by a magnetic flow meter (Foxboro 9300A) with an uncertainty of

±0.05 ft3/s.

An instrument carriage was mounted on horizontal steel rails and was moved along the

flume on wheels driven by a cable system and electric motor. Velocities and turbulence

quantities were measured with a 3D down-looking SonTek 10 MHz acoustic Doppler

velocimeter (ADV) as well as the 3D down-looking SonTek 16 MHz MicroADV. To measure

velocities and turbulence quantities near the piers and in the floodplain, a 2D and a 3D side-

looking SonTek MicroADV were utilized. The ADVs were attached to the instrument carriage

on a mobile point gage assembly that could be accurately positioned in all three spatial

dimensions. The 10 MHz ADV has a sampling volume of 0.015 in.3 while the sampling volume

of the MicroADV is approximately 0.005 in.3. The measuring volume is located 2.0 in. away

from the probe where velocities are measured based on the Doppler frequency shift of acoustic

signals reflected from small scattering particles moving at the same speed as the water. The

sampling frequency of the 10 MHz ADV was chosen to be 25 Hz, while a higher sampling

frequency of 50 Hz was possible with the MicroADV. A sampling duration of 2 minutes was

used at each measuring location.

Page 55: RP 2002 Final Report - Georgia

42

Velocities and turbulence quantities were measured over the temporarily fixed bed prior

to scour at relative heights above the bed of approximately 0.04, 0.1, 0.2, 0.4, and 0.6 times the

flow depth throughout the flow field including both near-field and far-field locations relative to

the bridge pier bent resulting in a total of approximately 750 measuring points for a single

experiment. In addition, detailed approach velocity profiles and turbulence quantities were

measured using a combination of the down-looking and side-looking ADVs.

Although Voulgaris and Trowbridge (1998) have demonstrated in flume experiments that

the ADV can measure both mean velocities and Reynolds stresses within 1 percent of the

measurements made by a laser Doppler velocimeter (LDV) and can describe the vertical

variation of Reynolds stresses according to accepted open channel flow results, the occurrence of

noise in measurements below a level of about 1.2 in. above the bed can cause problems. Some of

this noise is flow-related and can be attributed to high levels of both turbulence and mean

velocity shear near the bed. In addition, electronic noise can originate from errant reflections due

to the measuring volume being too close to the boundary and from boundary interference when

the return signal from the boundary interferes with the signal from the measuring volume (Lane

et al. 1998). One method of dealing with this noise is to filter the data according to the value of a

correlation coefficient that is a measure of the coherence of the return signals from two

successive acoustic pulses (Martin et al. 2002, Wahl 2002).

In this study it was found that the velocity and turbulence measurements near the bed

suffered from the same noise problems as experienced by other investigators, especially in the

near-field shear zone that experienced high turbulence levels. Accordingly, the data were filtered

by first requiring that the correlation coefficient of each sample in the 2-minute time record

exceed a value of 70 percent as recommended by the manufacturer (SonTek 2001) for obtaining

Page 56: RP 2002 Final Report - Georgia

43

turbulence statistics. In some cases, the filtering resulted in a large number of data points being

rejected from a given time record, so an entire two-minute record was rejected if the average

correlation coefficient fell below 70 percent.

Experimental Procedure

After completion of the flow field measurements over a fixed bed, the mobile sediment

bed was installed in the vicinity of the central pier bent, and scour experiments were conducted.

The flume was slowly filled to a depth larger than the test depth so as to prevent scour while the

test discharge was set. Then the tailgate was lowered to achieve the desired depth of flow.

Measurements of scour depth as a function of time at a fixed point were measured with the ADV

to determine when equilibrium had been reached. Then the flow rate was reduced while keeping

the scoured bed submerged, and the bed elevations were mapped in detail using the ADV feature

of acoustically pinging the bottom to measure the distance from the sampling volume to the bed.

Based on comparisons with point gage measurements, this method allowed the measurement of

bed elevation with an uncertainty of ±0.05 in. Some bed elevations very close to the pier were

measured directly with a point gage.

EXPERIMENTAL RESULTS

Summary of Data

Measured data for the Chattahoochee River bridge are presented in Table 3-2 as raw data and in

dimensionless form in Table 3-3. The raw data for the Flint River bridge are given in Table 3-4

with the data in dimensionless form summarized in Table 3-5. The raw data include the

measured discharge (Q), sediment size (d50), pier width (b), approach depth (y1), approach

velocity (V1), scour depth (ds), duration of scour experiment (T), and calculated scour

equilibrium time (Teq) according to the relationship developed by Melville and Chiew (1999).

Page 57: RP 2002 Final Report - Georgia

44

Table 3-2. Raw experimental data for Chattahoochee River bridge near Cornelia, GA. (b = pier width, y1 = approach flow depth, V1 = approach velocity, ds =scour depth, T = duration of scour, Teq = equilibrium time from Melville and Chiew, 1999).

1FB = flat-bed model 2RM = river model

Table 3-3. Dimensionless experimental data for Chattahoochee River bridge near Cornelia, GA (Fr = approach Froude number, Vc = critical velocity).

Run Model Scale Fr V1 /Vc y1 /b b/d50 ds /b No. Type Ratio 1 FB 0.043 0.350 0.64 3.62 13.9 1.23 2 FB 0.043 0.367 0.70 4.16 13.9 1.77 3 FB 0.043 0.511 0.91 3.32 13.9 2.91 4 FB 0.043 0.416 0.78 3.91 13.9 2.62 5 FB 0.043 0.404 0.74 3.72 13.9 2.17 6 FB 0.043 0.463 0.83 3.40 13.9 2.79 7 FB 0.043 0.399 0.76 4.16 13.9 2.25 8 FB 0.043 0.438 0.83 4.16 13.9 2.91 9 FB 0.043 0.602 1.07 3.32 13.9 2.93 1 RM 0.025 0.391 0.75 7.09 8.2 2.32 2 RM 0.025 0.429 0.82 7.09 8.2 2.33 3 RM 0.025 0.526 0.84 4.07 8.2 3.28 4 RM 0.025 0.557 0.88 3.95 8.2 2.96 5 RM 0.025 0.301 0.75 3.95 24.5 1.92 6 RM 0.025 0.334 0.83 3.95 24.5 2.22 7 RM 0.025 0.401 1.00 3.95 24.5 2.51 8 RM 0.025 0.232 0.71 7.04 24.5 2.22 9 RM 0.025 0.255 0.78 7.04 24.5 2.19

Scale Q d50 b Y1 V1 ds T Teq Run No.

Model Type Ratio (cfs) (mm) (ft) (ft) (ft/s) (ft) (hrs) (hrs)

1 FB1 0.043 6.92 3.3 0.151 0.546 1.468 0.185 24 25 2 FB 0.043 8.42 3.3 0.151 0.628 1.652 0.267 32 29 3 FB 0.043 8.42 3.3 0.151 0.502 2.054 0.440 36 37 4 FB 0.043 8.42 3.3 0.151 0.591 1.816 0.396 36 33 5 FB 0.043 8.42 3.3 0.151 0.561 1.717 0.328 36 31 6 FB 0.043 8.42 3.3 0.151 0.514 1.885 0.421 36 34 7 FB 0.043 9.63 3.3 0.151 0.628 1.793 0.339 36 32 8 FB 0.043 10.21 3.3 0.151 0.628 1.968 0.440 38 35 9 FB 0.043 10.21 3.3 0.151 0.502 2.422 0.443 37 41 1 RM2 0.025 4.50 3.3 0.089 0.628 1.760 0.205 48 20 2 RM 0.025 5.00 3.3 0.089 0.628 1.930 0.206 48 22 3 RM 0.025 2.45 3.3 0.089 0.360 1.790 0.290 48 23 4 RM 0.025 2.45 3.3 0.089 0.350 1.870 0.262 48 24 5 RM 0.025 1.35 1.1 0.089 0.350 1.010 0.170 48 32 6 RM 0.025 1.50 1.1 0.089 0.350 1.120 0.197 47 36 7 RM 0.025 1.66 1.1 0.089 0.350 1.347 0.222 47 41 8 RM 0.025 2.50 1.1 0.089 0.623 1.040 0.197 47 31 9 RM 0.025 3.00 1.1 0.089 0.623 1.143 0.194 47 34

Page 58: RP 2002 Final Report - Georgia

45

Table 3-4. Raw experimental data for Flint River bridge at Bainbridge, GA. (b = pier width, y1 = approach flow depth, V1 = approach velocity, ds =scour depth, T = duration of scour, Teq = equilibrium time from Melville and Chiew, 1999).

Scale Q d50 b Y1 V1 ds T Teq Run No.

Model Type Ratio (cfs) (mm) (ft) (ft) (ft/s) (ft) (hrs) (hrs)

1 FB1 0.020 7.71 3.3 0.120 0.50 1.708 0.164 48 26 2 FB 0.020 7.72 3.3 0.120 0.53 1.584 0.124 48 24 3 FB 0.020 8.85 3.3 0.120 0.52 1.762 0.175 48 27 1 RM2 0.011 1.50 1.1 0.068 0.500 0.810 0.071 48 17 2 RM 0.011 1.65 1.1 0.068 0.500 0.950 0.120 96 22 3 RM 0.011 1.80 1.1 0.068 0.500 1.100 0.190 96 27

1FB = flat-bed model 2RM = river model

Table 3-5. Dimensionless experimental data for Flint River bridge at Bainbridge, GA (Fr = approach Froude number, Vc = critical velocity).

Run Model Scale Fr V1 /Vc y1 /b b/d50 ds/b No. Type Ratio

1 FB1 0.020 0.426 0.75 4.17 11.1 1.37 2 FB 0.020 0.385 0.69 4.38 11.1 1.03 3 FB 0.020 0.431 0.77 4.33 11.1 1.46 1 RM2 0.011 0.202 0.57 7.35 18.8 1.04 2 RM 0.011 0.237 0.67 7.35 18.8 1.76 3 RM 0.011 0.274 0.78 7.35 18.8 2.79

1FB = flat-bed model 2RM = river model

The scour depths reported in Tables 3-2 and 3-4 are the maximum scour depths in front

of the most upstream pier in the main pier bent. Approach velocities and depths given in these

tables were measured upstream of the pier at a distance of 10 pier widths. The corresponding

dimensionless variables for each experimental run as identified in Eq. 3.1 are given in Tables 3-3

and 3-5. The critical velocity in the sediment mobility factor V1/Vc is calculated from Keulegan’s

equation with an equivalent sand-grain roughness of ks = 2d50. It is important to note that at the

same value of y1/b, a smaller sediment size produces a smaller Froude number and a larger value

of b/d50 for the same range of values of V1/Vc as shown in Table 3-3, for example, for Runs RM 1

and RM 8.

Page 59: RP 2002 Final Report - Georgia

46

The duration of the scour experiments is given in Tables 3-2 and 3-4 as well as the

calculated equilibrium scour time for maximum clear-water scour according to the relationship

developed by Melville and Chiew (1999). In general, the experimental scour durations are equal

to or greater than the equilibrium times. In fact, in most cases the scour experiments were

continued well beyond the estimated equilibrium time.

0.0

0.1

0.2

0.3

0.4

0 10 20 30 40 50

Time, hr

Scou

r dep

th,

ft

Exp. IExp. IIRegression

(a) Scour depth development with time in repeated experiments.

0

1

2

3

1.E+03 1.E+04 1.E+05 1.E+06

V c t /y 1

d s /b

Exp. IExp. IIRegression

(b) Dimensionless scour depth development with dimensionless time.

Figure 3.5. Scour development with time for Chattahoochee River model (Replicates of Run FB 8 with V1/Vc = 0.83 and y1/b = 4.2).

Page 60: RP 2002 Final Report - Georgia

47

The development of scour depth with time is shown in Fig. 3.5 (a) for two replicated

experiments in the Chattahoochee River flat-bed model (Run FB 8). The scour has clearly

reached equilibrium, and more importantly the data in Fig. 3.5 (a) show that the scour depth can

be reproduced in identical experiments with an uncertainty of approximately ±0.01 ft, or about

one grain diameter, if care is taken to level the bed to the same initial position. The scour

development with time is shown in dimensionless form in Fig. 3.5 (b), and it follows a linear

trend with the logarithm of dimensionless time as found by other investigators (Ettema 1980,

Sturm 1998, Cardoso and Bettes 1999).

Approach Velocity and Turbulence

The relative longitudinal and vertical turbulence intensities at the location of the approach

section to the main pier bent are shown in Fig. 3.6 for the Chattahoochee River model (Run FB

8) and in Fig. 3.7 for the Flint River model (Run RM 1) as typical examples. The longitudinal

and vertical turbulence intensities are shown in these figures nondimensionalized by the

longitudinal point velocity u as a function of relative height above the bed z/H where H =

approach flow depth. The experimental data are compared with accepted experimental

relationships for rough boundaries in open channel flow obtained by other investigators (Nezu

and Nakagawa 1995, Nikora and Goring 2000). Although there are some outliers close to the

bed, the data are within the experimental uncertainty of the accepted experimental relationships.

The comparisons shown in Figs. 3.6 and 3.7 are important not only because they verify the

acceptability of the turbulence measurements in the approach flow, but also because they provide

a baseline for comparison with turbulence intensities that exist in the near field of the pier bent

where scour occurs.

Page 61: RP 2002 Final Report - Georgia

48

Turbulence intensity in x-direction

0

0.1

0.2

0.3

0.4

0.5

0 0.2 0.4 0.6 0.8 1z/H

u'/u

10MHz ADVNezu and Nakagawa (1993)Nikora and Goring (2002)

Turbulence intensity in z-direction

0

0.1

0.2

0.3

0.4

0.5

0 0.2 0.4 0.6 0.8 1z/H

w'/u

10MHz ADVNezu and Nakagawa (1993)Nikora and Goring (2002)

Figure 3.6. Longitudinal and vertical relative turbulence intensity profiles at approach section for Chattahoochee River model (Run FB 8).

Turbulence intensity in x-direction

0

0.1

0.2

0.3

0.4

0.5

0 0.2 0.4 0.6 0.8 1z/H

u'/u

10MHz + 2D side ADVNezu and Nakagawa (1993)Nikora and Goring (2002)

Turbulence intensity in z-direction

0

0.1

0.2

0.3

0.4

0.5

0 0.2 0.4 0.6 0.8 1z/H

w'/u

10MHz + 2D side ADVNezu and Nakagawa (1993)Nikora and Goring (2002)

Figure 3.7. Longitudinal and vertical relative turbulence intensity profiles at approach section for Flint River model (Run RM 1).

Near-Field Turbulence Intensity

Turbulence intensities were measured for the flat-bed Chattahoochee River model at the 14

locations shown on the left and right sides of the main pier bent as shown in Fig. 3.8 (a) where

y/b is approximately ±2. For purposes of comparison, all turbulence intensities are

nondimensionalized by the mean approach velocity (U0) and are shown in Fig. 3.8 (b) for the

Page 62: RP 2002 Final Report - Georgia

49

(a) Measuring locations.

z/HLeft side of piersRight side of piers

0

0.1

0.2

0.3

0.4

0.5

0 0.5 10

0.1

0.2

0.3

0.4

0.5

0 0.5 10

0.1

0.2

0.3

0.4

0.5

0 0.5 10

0.1

0.2

0.3

0.4

0.5

0 0.5 10

0.1

0.2

0.3

0.4

0.5

0 0.5 10

0.1

0.2

0.3

0.4

0.5

0 0.5 1

u'/U

0

1 2 3 4 5 6

0

0.1

0.2

0.3

0.4

0.5

0 0.5 1

7

0

0.1

0.2

0.3

0.4

0.5

0 0.5 10

0.1

0.2

0.3

0.4

0.5

0 0.5 10

0.1

0.2

0.3

0.4

0.5

0 0.5 10

0.1

0.2

0.3

0.4

0.5

0 0.5 10

0.1

0.2

0.3

0.4

0.5

0 0.5 10

0.1

0.2

0.3

0.4

0.5

0 0.5 1

u'/U

0

8 9 10 11 12 13

0

0.1

0.2

0.3

0.4

0.5

0 0.5 1

14

(b) Longitudinal turbulence intensity relative to approach velocity U0.

z/HLeft side of piersRight side of piers

0

0.1

0.2

0.3

0.4

0.5

0 0.5 10

0.1

0.2

0.3

0.4

0.5

0 0.5 10

0.1

0.2

0.3

0.4

0.5

0 0.5 10

0.1

0.2

0.3

0.4

0.5

0 0.5 10

0.1

0.2

0.3

0.4

0.5

0 0.5 10

0.1

0.2

0.3

0.4

0.5

0 0.5 1

w'/U

0

1 2 3 4 5 6

0

0.1

0.2

0.3

0.4

0.5

0 0.5 1

7

0

0.1

0.2

0.3

0.4

0.5

0 0.5 10

0.1

0.2

0.3

0.4

0.5

0 0.5 10

0.1

0.2

0.3

0.4

0.5

0 0.5 10

0.1

0.2

0.3

0.4

0.5

0 0.5 10

0.1

0.2

0.3

0.4

0.5

0 0.5 10

0.1

0.2

0.3

0.4

0.5

0 0.5 1

w'/U

0

8 9 10 11 12 13

0

0.1

0.2

0.3

0.4

0.5

0 0.5 1

14

(c) Vertical turbulence intensity relative to approach velocity U0.

Figure 3.8. Near-field turbulence intensities for Chattahoochee River model (Run FB 8).

-10 -5 0 5 10 15 20

x/b

-5

0

5

y/b

U

1 148

Page 63: RP 2002 Final Report - Georgia

50

longitudinal direction and in Fig. 3.8 (b) for the vertical direction. The measured turbulence

intensities are close to the edge of the near field or wake zone and show a maximum near the bed

in absolute terms since the reference approach velocity U0 is a constant. The horizontal

turbulence intensity is roughly of the order of 10 percent of the mean approach velocity while the

vertical turbulence intensity is of the order of 5 percent of the mean approach velocity. Of

particular interest is the increase in turbulence intensities on the right side of the pier bent in

comparison to the left side for positions downstream of location 8 as defined in Fig. 3.8 (a). This

is the result of a slight skewness of the approach flow which is discussed in more detail in

Chapter 6.

Scour Contours and Near-Field Velocities

Fig. 3.9 shows the equilibrium scour contours for the same flat-bed Chattahoochee River model

run (FB 8) for which the turbulence intensities were given in Fig. 3.8. Superimposed on the scour

contours are the near-field velocity vectors measured after scour at a height above the bed of 40

percent of the depth.

Although the maximum scour depth for the experimental run shown in Fig. 3.9 is in front

of the upstream pier, scour holes are also evident in front of the downstream pier as well as

between the two inner piers underneath the solid web. Immediately downstream of the first and

fourth piers and underneath the inner pier web, a region of deposition, or less scour, can be

observed. This scour pattern will be explored in more detail in Chapter 6 in which the computed

three-dimensional velocity field is compared with the laboratory scour and velocity

measurements. The horizontal velocity vectors in Fig. 3.9 show the characteristic decrease in

magnitude associated with the approach to the pier stagnation line at a location of approximately

two pier widths upstream of the first pier. In addition, the velocity vectors at this location are at

Page 64: RP 2002 Final Report - Georgia

51

an angle to the centerline of the pier bent as the flow bends around the pier bent and separates to

form a wake zone. The horizontal velocities are very small between the piers as well as in the

downstream wake of the bent until the wake velocity defect gradually begins to recover in the

downstream direction.

-10 -5 0 5 10 15 20

x/b

-10

-5

0

5

10

y/b

ds/b

-3.0

-2.7

-2.4

-2.1

-1.8

-1.5

-1.2

-0.9

-0.6

-0.3

0.0

0.3

V/V1 = 1.0

Figure 3.9. Shaded scour contours and velocity vectors after scour at 40 percent of the approach depth for bank-full flow in the Chattahoochee flat-bed model (Run FB 8).

-10 -5 0 5 10 15 20

x/b

-10

-5

0

5

10

y/b

-3.0

-2.7

-2.4

-2.1

-1.8

-1.5

-1.2

-0.9

-0.6

-0.3

0.0

0.3

ds/b

V/V1 = 1.0

Figure 3.10. Shaded scour contours and velocity vectors after scour at 40 percent of the approach depth for bank-full flow in the Chattahoochee River model (Run RM 5).

Page 65: RP 2002 Final Report - Georgia

52

Scour contours and velocities after scour are shown for the Chattahoochee river model

case (Run RM 5) in Fig. 3.10 which is also for bank-full flow as in Fig. 3.9 for the flat-bed

model. The values of V1/Vc and b/d50 are not the same in Figs. 3.9 and 3.10, so only the scour

patterns and velocity vectors relative to the approach velocity are compared here while the scour

magnitudes will be discussed in the next section. The velocity field for the river model in Fig.

3.10 is not as uniform across the channel because of the river bathymetry, so the effect of the

obstruction caused by the pier bent on the near-field velocities is slightly different. In addition,

there is a greater skewness of the approach velocity with respect to the centerline of the pier bent

in the case of the river model. The angle of attack for the river model was measured to be 4.3°

while it was only 1.8° for the flat-bed model. This becomes apparent in the scour pattern in front

of the upstream pier in Fig. 3.10 in which the scour hole is offset to the right of the centerline of

the pier bent. Scour is less underneath the web between the inner piers in both Figs. 3.9 and 3.10,

and the continuity of the scour trench surrounding the pier bent is similar in both cases.

Scour patterns and velocity fields for the Flint River flat-bed model and river model are

given in Figs. 3.11 and 3.12, respectively. While there is an obvious scour hole in front of the

upstream pier, the deepest scour for the flat-bed model at bank-full flow (y1/b = 4.4) occurs in

two symmetric scour holes between the piers as shown in Fig. 3.11 (a). The velocity field after

scour in Fig. 3.11 (a) clearly shows the deceleration of flow as the upstream pier is approached

and the very small velocities between piers, but the horizontal velocity field does not explain the

scour pattern between piers. A closer look at the velocity components after scour in two cross

sections located just upstream of the first pier and between the two piers is given in Figs. 3.11 (b)

and 3.11 (c), respectively. The downflow induced by the first pier is still evident in Fig. 3.11 (b)

at the end of scour along with very small velocities near the bottom of the scour hole. However,

Page 66: RP 2002 Final Report - Georgia

53

-5 0 5 10 15

x/b

-5

0

5

y/b

-1.8

-1.5

-1.2

-0.9

-0.6

-0.3

-0.0

0.3

0.6

ds/b

A

A'

B

B' V/V1 = 1.0

(a) Scour contours and velocity vectors after scour at 40 percent of the depth.

6 4 2 0 2 4 6

y/b

-2

0

2

4

z/b

V/V1 = 0.24

Right sideLeft side

(b) Velocity vectors at cross section A-A’ after scour.

6 4 2 0 2 4 6

y/b

-2

0

2

4

z/b

V/V1 = 0.24

Right sideLeft side

(c)Velocity vectors at cross section B-B’ after scour.

Figure 3.11. Scour contours and velocity vectors after scour for bank-full flow in Flint River flat-bed model (Run FB 2).

Page 67: RP 2002 Final Report - Georgia

54

in Fig. 3.11 (c) there is an upwelling between the piers at the centerline of the pier bent with two

counter-rotating vortices on either side that coincide with the symmetric scour holes.

Scour patterns and the velocity field for flood flow (y1/b = 7.4) in the Flint River for the

river model are shown in Fig. 3.12. In this case, the symmetric scour holes disappear and the

deepest scour is in front of the upstream pier with scour of a lower magnitude between piers.

Skewness is a factor in this case since the angle of attack was measured to be 6.4° which in

combination with the river bathymetry helps to explain the small wake scour hole to the right of

the downstream pier and the deposition area also to the right and considerably downstream of the

pier bent. However, the most surprising difference in comparison with the flat-bed model can be

seen in Figs. 3.12 (b) and 3.12 (c) which show the velocities in cross sections upstream of the

first pier and between piers. There is an obvious secondary flow near the bed moving across the

cross section to the left which may originate from the floodplain flow joining the main channel

flow in the bridge opening or may be the result of the gradual bend in the river just upstream of

the bridge. In summary, there are significant three-dimensional flow effects in both the flat-bed

and river models, but they are due to different sources and they result in very different scour

patterns.

Comparison of Laboratory Scour Data with Prediction Formulas

To conclude this chapter on laboratory modeling, it is useful to compare the measured laboratory

scour depths at different model scales and for different sediment sizes with accepted clear-water

scour formulas. The scour depths being compared are at the nose of the upstream pier in the pier

bent. In the figures that follow, the measured dimensionless values of scour depth, ds/b, are

plotted together with predictions from scour formulas that include corrections for the

independent nondimensional variables listed in Eq. 3.1. The HEC-18 formula does not include

Page 68: RP 2002 Final Report - Georgia

55

-20 -15 -10 -5 0 5 10 15 20 25

x/b

-10

-5

0

5

10

y/b

-1.1-0.9-0.7-0.5-0.3-0.10.10.30.50.70.91.11.3

ds/b

V/V1 = 1.0

A

A'

B

B'

(a) Scour contours and velocity vectors after scour at 40 percent of the depth.

10 8 6 4 2 0 2 4 6 8 10

y/b

0

2

4

z/b

V/V1 = 0.5

Left side Right side

(b) Velocity vectors at cross section A-A’ after scour.

10 8 6 4 2 0 2 4 6 8 10

y/b

0

2

4

z/b

V/V1 = 0.5

Left side Right side

(c) Velocity vectors at cross section B-B’ after scour.

Figure 3.12. Scour contours and velocity vectors after scour for flood flow in Flint River model (Run RM 1).

Page 69: RP 2002 Final Report - Georgia

56

the effect of V1/Vc , while both the Sheppard and Melville formulas do not include the influence

of the Froude number. Accordingly, the plots of the dimensionless scour depths are given as a

function of V1/Vc with the value of the approach Froude number shown as a label on each data

point. Results are presented in separate figures for the flat-bed model and the river model

because the values of b/d50 are different in the two cases, and they have a significant effect in the

Sheppard and Melville formulas (reduction in predicted scour for values of b/d50 less than about

25). On the other hand, the value of b/d50 is not a factor in the HEC-18 formula. Finally, both

bank-full and flood flow experimental runs are included in the figures with values of y1/b

approximately equal to 4.0 and 7.0, respectively, for both the Chattahoochee River and the Flint

River. These values of y1/b are large enough that they have almost no influence in the Sheppard

and Melville formulas, while the value of y1/b appears explicitly in the HEC-18 formula

regardless of its value.

The comparison between measured and predicted scour depths for the flat-bed

Chattahoochee model (1:23.3 scale) at bank-full flow is shown in Fig. 3.13. The data points at

low values of V1/Vc are bounded above by the Sheppard and Melville formulas. As V1/Vc

continues to increase, the data points approach the HEC-18 formula which is considerably above

the other two formulas for this case. The values of the Froude number also become rather large at

the higher values of V1/Vc. This is a consequence of the larger model sediment size (3.3 mm)

which also produces smaller values of b/d50. These values of the Froude number are considerably

higher than the field value for bank-full flow which is approximately 0.33.

Figure 3.14 shows the scour depth comparisons for the river model of the Chattahoochee

River (1:40 scale). The sediment size for these data is 1.1 mm which results in a value of b/d50 of

Page 70: RP 2002 Final Report - Georgia

57

0.0

1.0

2.0

3.0

4.0

5.0

0.5 0.6 0.7 0.8 0.9 1.0 1.1V 1 /V c

d s/ b

Flatbed model, b/d50=13.9Sheppard (2003), b/d50=13.9Melville (1997) , b/d50=13.9HEC-18(1995), b/d50=13.9

0.35

Fr=0.600.510.44

0.46

0.42

0.400.40

0.37

Figure 3.13. Comparison of measured scour depths in Chattahoochee River flat-bed model with scour prediction formulas for bank-full flow (y1/b = 4.0).

0.0

1.0

2.0

3.0

4.0

5.0

0.5 0.6 0.7 0.8 0.9 1.0 1.1V 1 /V c

d s/ b

River model, b/d50=24.5Sheppard (2003), b/d50=24.5Melville (1997) , b/d50=24.5HEC-18 (1995), b/d50=24.5, y1/b=7.0HEC-18 (1995), b/d50=24.5, y1/b=4.0

0.300.23 0.33

Fr=0.40.26

Figure 3.14. Comparison of measured scour depths in Chattahoochee River model with scour prediction formulas for bank-full flow and flood flow (y1/b = 4.0 and 7.0).

Page 71: RP 2002 Final Report - Georgia

58

24.5 at which there is almost no influence on predicted scour depths in the Sheppard and

Melville formulas. Another result of the smaller sediment size is a range in the laboratory Froude

number values that is smaller and more comparable to the prototype value. There is good

agreement between the measured data and all three scour prediction formulas in Fig. 3.14.

The comparison between measured and predicted scour depths for the flat-bed Flint River

model (1:50 scale) at bank-full flow is shown in Fig. 3.15. As for the Chattahoochee model

results in Fig. 3.13, the Sheppard and Melville formulas agree with the measured scour depths at

small values of V1/Vc, but the corresponding HEC-18 formula predictions are well above the

measured values because of the small values of b/d50 for the 3.3 mm sediment size. The

difference in this case is that measurements were not made at very high values of V1/Vc so that

the Froude numbers are not excessively large.

Fig. 3.16 shows the scour depth comparisons for the Flint River model (1:90 scale)

utilizing the 1.1 mm sediment at flood flow stage (y1/b = 7). As in Fig. 3.14 for the

Chattahoochee River model, the Sheppard and Melville scour prediction formulas provide

reasonable predictions of the scour depth, while the HEC-18 formula overpredicts the scour

depth somewhat more than in Fig. 3.14 because the model value of b/d50 is less than 25. The

Froude numbers in Fig. 3.16 for the Flint River model remain relatively low and comparable

with the field value of 0.25.

Further comparisons of laboratory scour depths with field values are made in Chapter 6.

The results of this chapter indicate that good comparisons with the Melville and Sheppard

formulas can be obtained even at different model scales if the model sediment size is small

enough that values of the Froude number do not become too large. The HEC-18 formula, on the

other hand, tends to overpredict the scour depth for most of the measurements that were made.

Page 72: RP 2002 Final Report - Georgia

59

0.0

1.0

2.0

3.0

4.0

5.0

0.5 0.6 0.7 0.8 0.9 1.0V 1 /V c

d s/ b

Flatbed model, b/d50=11.1Sheppard (2003), b/d50=11.1Melville (1997) , b/d50=11.1HEC-18(1995), b/d50=11.1

Fr=0.3

0.430.43

Figure 3.15. Comparison of measured scour depths in Flint River flat-bed model with scour prediction formulas for bank-full flow (y1/b = 4.0).

0.0

1.0

2.0

3.0

4.0

5.0

0.5 0.6 0.7 0.8 0.9 1.0V 1 /V c

d s/ b

River model, b/d50=18.8Sheppard (2003), b/d50=18.8Melville (1997) , b/d50=18.8HEC-18 (1995), b/d50=18.8

0.20

Fr=0.27

0.24

Figure 3.16. Comparison of measured scour depths in Flint River model with scour prediction formulas for flood flow (y1/b = 7.0).

Page 73: RP 2002 Final Report - Georgia

60

CHAPTER 4. THREE-DIMENSIONAL NUMERICAL MODELING

INTRODUCTION

The flow features around bridge foundations on natural river reaches are largely determined by

the geometry of the structures, and by the complex bathymetry of the river. There is a large

disparity between the chacteristic length scales of the river flow and those produced by the

bridge foundation. The existence of bridge piers and abutments in the flow introduces vortex

shedding from the concrete walls to the already highly three-dimensional and strongly turbulent

flow existing in the river.

To simulate the flow around bridge foundations, a novel state-of-the-art numerical solver

has been developed which has the following features:

1) It resolves accurately and efficiently the intricate geometric details of real-life bridge

foundations, which typically consist of multiple bundles of piers with complex geometry;

2) It accounts for the large-scale topography of the river reach where the bridge is

embedded;

3) It captures the large-scale dynamics of the coherent vortex shedding, modeling the small

scale turbulence with statistical turbulence models.

The numerical model simulates the turbulent flow by solving the unsteady Reynolds-averaged

Navier-Stokes (URANS) equations along with the standard k-ε model. Overset or Chimera grid

techniques are employed in the calculations to deal with the arbitrarily complex geometry of the

problem. The following sections provide detailed descriptions of the numerical model that was

developed as well as the results that were obtained from the numerical simulations.

Page 74: RP 2002 Final Report - Georgia

61

NUMERICAL METHOD

The 3D, incompressible, URANS and turbulence closure equations are formulated and solved in

generalized, curvilinear coordinates in strong conservation form. The governing equations are

discretized in space using the three-point, central, second-order accurate, finite-volume scheme.

Third-order, fourth-difference, matrix-valued artificial dissipation terms (Lin and Sotiropoulos

1997) are explicitly added to the discrete equations to suppress grid-scale oscillations. The

discrete equations are integrated in time using a second-order-accurate, dual- or pseudo-time-

stepping, artificial compressibility scheme. The equations are integrated in pseudo time using the

Beam-and-Warming approximate-factorization scheme in conjunction with V-cycle multigrid

and local-pseudo-time-stepping for faster convergence. At the inlet, fully-developed turbulent

flow distributions for the mean-velocity field and the turbulence quantities are specified. Non-

reflecting, characteristic-based boundary conditions are applied at the outlet boundary to allow

vortical structures to exit the flow domain without distortion. The free surface is treated as a flat,

rigid lid. Wall functions are used to account for the wall roughness and flow turbulence near all

solid walls. The developed numerical solver is parallelized to take full advantage of modern

computational power. Typically, 15 pseudo-iterations are required to reduce all residuals by three

orders of magnitude, and a total number of 3000 physical time steps are required to obtain a

statistically converged solution.

GRID GENERATION FOR COMPLEX BATHYMETRY AND OBSTRUCTIONS

An example of the complex geometries of interests is given in Fig. 4.1, which shows the actual

river bathymetry of a reach of the Chattahoochee River near Cornelia, GA with a single bent of

bridge piers mounted on the bed. The bent consists of four rectangular piers located one behind

the other along the flow direction. There is a web connecting the two middle piers which does

Page 75: RP 2002 Final Report - Georgia

62

not extend all the way to the channel bed. The actual bridge foundation consists of three such

bents across the bridge span. In this study, the focus is on the single-bent case. In the following

section, the overset grid method is described in the context of the flow around the bridge

foundation of the Chattahoochee River case.

Figure 4.1. Numerical geometry of a single bent of piers on natural river reach.

(a) (b)

Figure 4.2 Overset grid system.

Page 76: RP 2002 Final Report - Georgia

63

The idea of the overset grid method is to divide the original complex flow domain into

several small subdomains which are simple for grid generation. These subdomains are allowed to

overlap with other subdomains. The overset grid layout used for the site on the Chattahoochee

River near Cornelia, GA is illustrated in Fig. 4.2. As shown in Fig. 4.2(a), a curvilinear grid

system, which will be referred to as the background grid hereafter, is used to discretize the river

reach---denoted as Subdomain 1 in Fig. 4.2(a). Other four subdomains, Subdomain 3, 4, 5, and 6

as shown in Fig. 4.2(a), are employed to discretize the four bridge piers mounted on the river bed

and the web connecting the two middle piers. Such a grid layout allows, on the one hand,

resolution of all essential flow features resulting from the river bathymetry with the background

grid, while the grids used in the other subdomains make it possible to cluster grid nodes in the

vicinity of solid walls, where the majority of the unsteady flow physics occur. Due to the

disparity in spatial scales between the river reach and the pier bent, there is a large discontinuity

between the grid spacing of the background grid (Subdomain 1) and that of Subdomains 3, 4, 5,

and 6 used to discretize the piers. As shown in Tang et al. (2003), large discontinuities in mesh

spacing across subdomain interfaces tend to introduce spurious oscillations in the computed flow

variables and in general lead to inaccurate solutions. In order to remedy this problem, yet another

subdomain, Subdomain 2 in Fig. 4.2(a), is embedded between the background grid and the pier-

bent grids. Grid nodes are distributed uniformly in this subdomain and their total number is

selected such that the spatial resolution of this grid alleviates the large discontinuity between the

coarse resolution of the background grid and the very fine resolution of the pier-bent grids. Grid

embedding with Chimera overset grids allows both the large-scale flow features controlled by

the river topography and the complex unsteady flow induced by the piers to be taken into

account without requiring excessively large grid sizes.

Page 77: RP 2002 Final Report - Georgia

64

The solutions on the original complex domain are obtained in an iterative manner.

Governing equations are first solved separately in each subdomain and the solutions

commnunicate with the neighboring subdomains through interpolation on the subdomain

interfaces. The interface interpolation is implemented in two steps. First, the location of each

grid node on the interface relative to the other domains is determined using a Newton method,

which is an iterative method to obtain a numerical solution of a system of equations of the form

f(x) = 0. This grid connectivity information can be performed as a preprocessing step and stored

for use during the calculation. The second step is the interpolation of the flow variables from

Subdomain 1 (which contains the background flow field information) to the interface of interest.

A second-order interpolation scheme is used to maintain the overall solver accuracy. The code

developed for this study features both standard trilinear interpolation for all flow variables at grid

interfaces (Steger and Benek 1987) as well as the so-called mass-flux based interpolation

approach developed in Tang et al. (2003).

APPLICATION OF THE 3D MODEL

The developed numerical solver was applied to study the 3D unsteady turbulence flows of the

following cases. In all these cases, the actual geometry of the bridge foundations was used. They

were either mounted on the actual river bathymetry or on flat river beds:

1) Bridge in the Chattahoochee River near Cornelia, GA:

a. Single pier bent mounted on the actual river bed;

b. Single pier bent mounted on a flat river bed;

c. Two pier bents mounted on a flat river bed.

2) Single pier bent of the Flint River foundation mounted on a flat river bed.

3) Single pier bent of the Ocmulgee River foundation in Macon, GA, mounted on a flat bed.

Page 78: RP 2002 Final Report - Georgia

65

Only results from Case 1 are presented in this section in order to demonstrate the general

capabilities of the numerical model and underscore the complexity of hydrodynamics in the

vicinity of real-life bridge foundations. Results from all three cases will be presented in Chapter

6 in the numerical model validation section.

The grid shown in Fig. 4.1 is used to study the flow through the Chattahoochee River

reach. The calculations are carried out for flow parameters that are typical for the specific reach,

where the bulk velocity is about 7 ft/s.

First, the capacity of the developed numerical solver for resolving the flow unsteadiness

in the vicinity of bridge foundations is illustrated. In Fig. 4.3, the calculated time histories of the

velocity component along the y axis at two points in the flow are shown. Point A is located in

Subdomain 6 just downstream of the last pier while Point B is located in the background grid

(Subdomain 1)---for point locations (see Fig. 4.2). As seen in Fig. 4.3, following an initial

transient of approximately 200 time units, the flow at Point A attains an oscillatory, periodic

state. In stark contrast, the flow at Point B reaches a quasi-steady state after the initial transition--

-a very weak unsteady fluctuation persists at this point but its amplitude is only a very small

fraction of the mean vertical velocity at Point B. This is consistent with the fact that a) the

location of point B is far away from the bridge foundations; and b) the grid resolution in the

background region is really coarse and designed to provide the background flow features for the

bridge foundations. Note that this flow unsteadiness was captured under steady inflow velocity

conditions. The flow unsteadiness is excited naturally by the large-scale vortex shedding induced

by the bridge foundations.

Page 79: RP 2002 Final Report - Georgia

66

Figure 4.3. Dimensionless velocity-time history at two different points A and B (U0 = approach velocity; T = b/U0; b = pier width).

To illustrate the complexity of the large-scale flow in the vicinity of the piers, Fig. 4.4

shows two snapshots of instantaneous velocity magnitude contours and three-dimensional

streamlines at the corresponding instants of time. This figure illustrates clearly the complexity of

the flow in the vicinity of the piers. Features such as the unsteady meandering of the shear layers

around the bent and the unsteadiness of the recirculating flow regions in the pier wakes are

clearly evident in these results. The effect of the complex river bathymetry on the pier hydraulics

is also apparent. Note that for both instants in time, the instantaneous velocity contours are

highly asymmetric. The velocities are considerably higher and the vortex shedding is more

intense on the left side of the pier bent (as viewed looking downstream) due to the fact that the

approach flow is skewed by about 5° relative to the streamwise axis of the bent. Obviously such

complex flow patterns cannot be accurately simulated without taking into account the complexity

of the ambient bathymetry. As shown in Fig. 4.4(b), highly unsteady, coherent vortical structures

with axes perpendicular and parallel to the bed are seen to appear and disappear continuously

throughout one period of the flow.

Page 80: RP 2002 Final Report - Georgia

67

(a) Velocity contours. (b) 3D streamlines.

Figure 4.4. Visualization of instantaneous flow field in the vicinity of bridge piers.

The above results clearly reveal the existence of large-scale coherent vortex structures in

the vicinity of the bridge foundations. In order to further understand the complex flow physics in

these regions, the flow is studied by replacing the complex river bed bathymetry with a flat bed

in cases 1(b), 2, and 3—for this configuration laboratory experiments were carried out and the

data from these experiments are used to validate the numerical model in Chapter 6 of this report.

The incoming flow is assumed to be aligned with the axis of the bridge piers. The same incoming

flow velocity as in the laboratory experiments is used for the numerical simulations.

Page 81: RP 2002 Final Report - Georgia

68

The grid system used for this calculation is shown in Fig. 4.5(a). As illustrated in Fig.

4.5(b), where the time history of spanwise velocity components at two different points located on

the symmetric plane of the pier-bent axis is shown, the flow unsteadiness is captured. After an

initial transient stage, the flow approaches a periodic state, and this flow unsteadiness is

sustained throughout a long calculation time span. Juxtaposing the two snapshots of resolved

streamwise velocity contours shown in Fig. 4.6(a) with the time-averaged flow at the horizontal

plane shown in Fig. 4.6(b) clearly shows that unsteadiness in the flow originates due to the

Kelvin-Helmoltz type instability of the shear layers emanating from the upstream corner of the

foundation and the intense vortex shedding in the wake. The transverse flapping and meandering

of the wake flow is clearly evident in the two snapshots in this figure. Furthermore, it should be

noted that the time-averaged flow field shown in Fig. 4.6(b) exhibits a high degree of symmetry

with respect to the streamwise axis of the foundation, thus suggesting that the simulated time

interval is suffcient for obtaining statistically-converged mean flow.

Page 82: RP 2002 Final Report - Georgia

69

(a) Overset grid layout. (b) Time history of spanwise velocity.

(c) Measurement and comparison locations.

Figure 4.5. Bridger piers mounted on flat bed.

Page 83: RP 2002 Final Report - Georgia

70

(a) Streamwise velocity contours at two different time instants on a horizontal plane.

(b) Time-averaged streamwise velocity contours.

Figure 4.6. Streamwise velocity contours.

Figure 4.7. Snapshot of instantaneous streamlines of the large-scale flow.

Page 84: RP 2002 Final Report - Georgia

71

Fig. 4.7 shows a snapshot of streamlines in the vicinity of the bridge piers. The complex

coherent vortex structures are clearly reproduced in this figure. In front of the first bridge pier on

the upstream side, the flow is smooth and organized with the horse-shoe vortex forming in the

area of conjunction between the downward flow and the riverbed. The flow behind the first

bridge pier is extremely complex, dominated by multiple tornado-like vortices. These vortices

are the result of the wake flow of the front pier and the solid wall of the following bridge piers.

The snapshot clearly illustrates the result of the interaction of the flow and the bridge

foundations, and these complex vortex structures are anticipated to play an important role in the

formation of the local scour hole.

Another case study to be presented is the flow around two bents of bridge piers mounted

on a flat riverbed. The geometries of the piers and the distance between these two bents are based

on the site near Cornelia, GA. The streamwise velocity contours on a horizontal plane just below

the water surface are shown in Fig. 4.8. As seen in this figure, despite the relatively large

distance between the two bents, there appear to be very weak interactions of the vortical

structures emanating from the solid walls of the piers. To quantify the interaction, the total

turbulence kinetic energy obtained from the numerical solution is calculated. Fig. 4.9 shows the

turbulence kinetic energy profile along the depth direction at four locations on both sides of the

right-hand pier bent (see Fig. 4.8 for locations). These locations are symmetric about the

symmetry axis of the bent. If the flow between these two bent structures has very weak or no

interaction, these profiles are anticipated to be symmetric on both sides. However, as shown in

Fig. 4.9, they are asymmetric suggesting that the turbulence structure in the vicinity of the two

bents is affected considerably by their mutual interaction in spite of their relatively large spacing.

Page 85: RP 2002 Final Report - Georgia

72

Figure 4.8. Instantaneous streamwise velocity contours.

Figure 4.9 Turbulence kinetic energy profile (see Figure 4.5 (c) for locations).

Page 86: RP 2002 Final Report - Georgia

73

CONCLUSIONS

The results presented in this chapter point to the following conclusions:

1) The flows around natural river bridge foundations are naturally unstable and highly three-

dimensional. Large-scale coherent vortex shedding exists in the vicinity of the bridge

foundations with multiple vortices having axes both parallel and perpendicular to the bed;

2) Both the river bathymetry and the presence of multiple pier bents can influence the flow

patterns considerably and need to be taken into account if realistic flow predictions are to be

obtained.

3) The numerical model developed in this research is a powerful engineering simulation tool for

elucidating the complex flow physics of real-life bridge foundation flows. As will

subsequently be shown, the insights into the flow structures provided by the numerical model

facilitate the understanding and interpretation of the results from laboratory experiments for

foundations mounted on deformable beds.

Page 87: RP 2002 Final Report - Georgia

74

CHAPTER 5. SEDIMENT EROSION PROPERTIES

INTRODUCTION

Shelby tube sediment samples collected from the foundations of ten (10) bridges located in the

state of Georgia were tested in the laboratory to find their erosional behavior and the correlation

of erosion parameters with sediment properties in order to improve the prediction of scour

around bridge foundations. These sites were spatially distributed in order to fall into different

major river basins and in different physiographic regions. Flume measurements were performed

using a rectangular, tilting, recirculating flume that was modified to receive Shelby tube samples

in the flume bottom that could be extruded upward as erosion occurred. Regression analysis was

performed on erosion rates as a function of applied shear stress to determine the parameters of

the erosion function. The resulting parameters, the critical shear stress and the erosion rate

constant, were correlated with soil properties and physiographic regions, and the results are

reported in this chapter. Additional details can be found in the thesis by Navarro (2004).

SEDIMENT EROSION RESISTANCE

The size of the particles being eroded is the principal factor that influences the type of forces

involved in resisting erosion. For coarse sediments with low content of fine particles (smaller

than 76 microns) gravity forces govern the process. Alternatively, when fine sizes are present in

the bed material, additional forces become important. For instance, for particle sizes smaller

than 10 microns or clay sizes, electrical forces make their appearance. In addition, the particle-

fluid interaction cannot be disregarded. The ionic concentration and pH of the fluid affect

particle charges; therefore the erosion process becomes dependent on the chemistry of the pore

water (Ravisangar et al. 2001).

Page 88: RP 2002 Final Report - Georgia

75

As the particles get smaller, electrical forces have more effect on the erosion resistance

because electrical forces are highly dependent on the particle’s specific surface area, which is

defined as the surface area per unit volume or mass of the particle. Given that specific surface

area increases as the size decreases, and that it is greater for platy particles than for spherical

particles, an abrupt change in the behavior of the forces takes place as the size changes from silty

to clay-size material. Silty material is the result of mechanical weathering and therefore

maintains a rounded shape. On the other hand, clay has platy structures with high values of

specific surface area. As a result, the interparticle forces that resist the hydrodynamic drag force

and erosion include the gravitational force, Coulombian attraction, van der Waals attraction and

double layer repulsion. Short range forces, such as hydration forces and Born repulsion, may

also be important in determining the net overall attractive or repulsive force between clay

particles (Mahmood et al. 2001).

The gravitational force manifests itself as the submerged weight of the sediment particle,

and it is the only force resisting erosion for coarse-grained sediments. Erosion is caused by the

hydrodynamic force consisting of lift and drag components produced by a viscous fluid moving

around a particle. Both surface drag and form drag contribute to the total drag force. The

movement of coarse-grained particles is produced when the hydrodynamic force overcomes the

submerged weight. The hydrodynamic force in the sediment transport literature is often

represented simply by the applied shear stress (τ), or force per unit surface area. The applied

shear stress for steady uniform open channel flow can be measured at the threshold of movement

of sediment particles on the channel bed and defined as the critical shear stress, τc which can be

given in dimensionless form as (Sturm 2001)

⋅==

⋅−=

νρτ

γγτ

τd

fd

cc

ws

cc

2/1

*2*)/(

Re)(

(5.1)

Page 89: RP 2002 Final Report - Georgia

76

in which γs - γw = submerged specific weight; d = particle diameter; ρ = water density; ν =

kinematic viscosity; τ*c = Shields Parameter; and Re*c = critical boundary or particle Reynolds

number. These last two parameters were presented by Shields for the initiation of sediment

motion in the widely known Shields Diagram. The first parameter (τ*c) can be interpreted as the

ratio of the hydrodynamic force per unit area to the gravitational force per unit volume. The

second parameter (Re*c) represents the ratio of the particle diameter to the thickness of the

viscous sublayer. In the definition of the last two parameters the diameter and the critical shear

stress are included, which impedes the direct calculation of the critical shear stress for a given

diameter. The introduction of a third dimensionless parameter, given by [0.1Re*c2/τ*c]1/2,

eliminates this restriction. The result of that analysis can be given in the form (Julien 1995)

⋅⋅−==

3/1

2

3

*3*)1(

ντ dgSGdfc (5.2)

in which d* = dimensionless particle diameter and SG = specific gravity. The functional

relationship suggested by Eq. 5.2 has been developed experimentally by many investigators for

coarse-grained sediments and it is shown in Fig. 5.1.

Page 90: RP 2002 Final Report - Georgia

77

0.01

0.1

1

0.1 1 10 100 1000Dimensionless Diameter, d *

Shie

lds P

aram

eter

, τ*c

Figure 5.1. Shields diagram for direct determination of critical shear stress of coarse-grained sediments (after Sturm 2001).

For fine-grained materials, in addition to gravitational forces opposing the movement,

interparticle forces start to act. Among these can be mentioned Coulombian attraction, van der

Waals attraction, and double layer repulsion. Coulombian attraction acts when there are counter

electrical charges interacting. Edges of clay particles that have a positive charge are attracted to

the negative face of the mineral. Van der Waals attraction is a result of the nonuniform charge

distribution on adjacent molecules. The closer the particles can move towards each other, the

stronger that this force will be. Its value depends on the shape of the interacting particles. The

double layer repulsion force makes erosion more likely. This force acts to cause particle

repulsion, making erosion by hydrodynamic forces easier. It depends on the ionic concentration

of the environment more than any other factor. As a result of this dependence, when ionic

concentration changes, this force changes and affects the equilibrium of the system. Coulombian

Page 91: RP 2002 Final Report - Georgia

78

and van der Waals forces will redistribute to form a new equilibrium state with a new particle

arrangement.

The relative contribution of the interparticle forces depends on the size and structure of

sediments. For example, considering two relevant forces, gravitational and van der Waals, it can

be shown that the gravity force is overwhelmed by the van der Waals force for particles smaller

than 60 microns (Santamarina 2001). Thus, the applied hydrodynamic drag force, which is a

combination of surface and form drag as well as turbulent bursts near the bed, must overcome

not only the gravitational forces but also the much larger interparticle forces for fine-grained

sediments.

The arrangement of the sediment particles is an important determinant of erosion.

Sediments that form a well-arranged structure will have great erosion resistance than sediments

that do not. The arrangement of particles depends on pH and ionic concentration conditions

present when they are deposited. Fine-grained platy particles have three main associations:

Edge-to-Face, Edge-to-Edge, and Face-to-Face. Edge-to-Face (E-F) arrangements are governed

by Coulombian forces presented by the contrary charge of the faces and edges of the particles.

Face-to-Face (F-F) arrangements are present when sedimentation occurs at high concentrations

and the van der Waals force prevails over double layer repulsion. Edge-to-Edge (E-E) are

transition arrangements of the other two structures given that there is not a dominant force that

governs the final structure.

Ravisangar et al. (2001) analyzed the influence of sediment pH on bed structure

formation and on initial erosion rates in flume experiments on kaolinite sediment. For different

pH conditions, kaolinite sediments have different bed structures and therefore different values of

bulk density and water content. Initial erosion rates were measured for those structures with the

Page 92: RP 2002 Final Report - Georgia

79

following results. For pH conditions below 5.5, Edge-to-Face associations predominate. As the

pH increases, the bed structure becomes weaker, corresponding to Edge-to-Edge associations

dominating the structure. At pH conditions above 7, the sediment particles associate as Face-to-

Face increasing resistance to erosion. In terms of erosion resistance, the strongest associations

were E-F and F-F, while erosional strength was less in the transition stage.

Mahmood et al. (2001) performed experiments on attachment and detachment of particles

in porous media columns and calculated interparticle forces for platy particles including van der

Waals forces, electrical double layer forces, hydration forces and Born repulsion. An

interparticle force model was developed using the actual shape of kaolinite particles (hexagonal

platelet-like), and it produced results consistent with experimental observations. The force

magnitudes followed the decreasing sequence F-F > E-F > E-E, which is the same sequence

observed by Ravisangar et al. (2001) in flume erosion experiments for kaolinite. In addition, the

pH value observed for the maximum in the percent detachment was around 5, where E-E

associations predominate.

Because most of the sediments sampled at bridge foundations in this project consisted of

some fraction of fine-grained sediments, the interparticle forces just discussed are an important

consideration in determining erosion resistance. Despite the advances that have been made in

calculation of interparticle forces, experimental evaluation is still the only viable alternative to

estimating erosion resistance. As a result, flume experiments were conducted to evaluate the

relative erosion resistance of the sediment samples.

EROSION MEASUREMENT

The most important measures of erosion are the critical shear stress and the erosion rate of the

sediment when it is resuspended and transported at higher shear stresses. Different types of

Page 93: RP 2002 Final Report - Georgia

80

experimental equipment have been used to measure erosion properties. These include the linear

recirculating flume, rotating annular flume and submerged impinging jet, among other devices.

Briaud et al. (1999), developed an apparatus called the erosion function apparatus (EFA), which

can test Shelby tube samples by introducing them into a rectangular duct having a cross section

50.8 mm high and 101.6 mm wide. In addition to laboratory measurements, in situ erosion

measurements have also been attempted as in the case of Ravens and Gschwend (1999) who

performed measurements of sediment erodibility in Boston Harbor. Still, the most common

devices to study sediment erosion phenomena experimentally are the linear recirculating flume

and the rotating annular flume.

The linear recirculating flume (Partheniades 1965; McNeil et al. 1996; Dennett et al.

1998; Ravisangar et al. 2001) used in this study is basically a straight open channel with an open

section at the bottom through which a sample of the erodible material is introduced. The flow

conditions in the channel are adjusted in order to assure fully developed, uniform, turbulent flow

as well as to apply a known shear stress at the bed. A piston is used to extrude the sample into

the flume as it is eroded. The height of the material eroded is recorded continuously as well as

the time during which it occurs, which when multiplied by the cross sectional area of the sample

results in the volumetric or gravimetric, if sediment density is known, erosion rate corresponding

to the specified flow conditions. A similar procedure is followed by Briaud et al. (1999), with

the difference that the flow through the sample is pressurized.

EROSION RELATIONSHIPS

Many attempts have been made to relate erodibility to bulk variables of the sediment and pore

fluid, and to the conditions of the flow. Among these contributors are McNeil et al. (1996), who

made measurements of erosion of undisturbed bottom sediments. Since erosion properties vary

Page 94: RP 2002 Final Report - Georgia

81

spatially throughout a nonhomogeneous material, they were measured as a function of depth.

McNeil et al. (1996) found the most important sediment parameters that affect the erosion

phenomenon are the bulk density, water content, average particle size, and organic content.

Among the fluid characteristics that affect the erodibility, pH was already mentioned for the

results of Ravisangar et al. (2001). It was found that the erosion resistance depends on the bed

structure, which in turn is dependent on the pH of the pore water.

Zreik et al. (1998) and Hoepner (2001) attempted to relate erosional behavior to more

conventional measures of soil strength. Zreik et al. (1998) compared the erosional and

mechanical strength of deposited fine sediment. Mechanical strength was measured in a manner

similar to the conventional fall cone, where a cone is released from the sediment surface and

penetrates by its own weight for a period of time. Their results showed that erosional strength

was one order of magnitude smaller than mechanical strength. Two hypotheses are presented by

Zreik et al. (1998): first, the resistance to erosion is governed by the weakest of the individual

bonds between flocs, while mechanical resistance is governed by the group of bonds between the

flocs available in the sheared sediment mass. Second, for the erosional phenomenon, turbulent

eddies in the flow accumulate greater energy that causes sporadic motion of individual flocs

before the bulk shear strength of the bed is mobilized. Hoepner (2001) related the stability of

fine sediments tested in flume experiments to rheometer measures of yield stress and found that

the measured yield stress can be a practical index to predict the erosional strength of undisturbed

sediment.

Briaud et al. (2001) using their EFA (Erosion Function Apparatus) measured the erosion

rate of fine grained soils, finding that the most common shape of the erosion rate vs. applied

shear stress curve is concave up. However, straight and convex shapes were also found. The

Page 95: RP 2002 Final Report - Georgia

82

convex shape was associated with the change of mechanism from surface to mass erosion.

Briaud et al. (2001) also correlated the erosion function with soil properties. One of the curve-

fitting parameters involved, the critical shear stress, is thought to increase when the soil unit

weight, plasticity index, soil shear strength, or fines content increase; and to decrease when the

void ratio, soil swell, dispersion ratio, soil temperature or water temperature decrease. However,

poor correlations were found with the plasticity index, undrained shear strength, and percent

passing the #200 sieve, for example. On the other hand, the initial slope of the erosion rate vs.

applied shear stress curve showed an encouraging relationship with the critical shear stress.

The dissimilar approaches to finding a unique relationship for the erosion resistance of

sediments is due to the difficulty in characterizing the microstructure based on the macro

properties of the material. This is in particular difficult for fine-grained sediments for which the

Shields relationship does not apply. Microstructure properties such as interparticle distance or

particle arrangement are not easily converted into particle size distribution and bulk density. In

addition, the nonhomogeneity of natural sediments adds more uncertainty in the measured

sediment properties. These problems force the use of experimental work to measure soil

erodibility although an analytical approach can provide a better understanding of the

phenomenon.

EXPERIMENTAL METHODS

Sample Locations and Properties

The Georgia Department of Transportation (GDOT) supplied the sediment core samples tested in

this project. The flume experiments were conducted on material collected according to ASTM

(D 1587-00): Standard practice for thin-walled tube sampling for geotechnical purposes. In this

project, thin-walled tubes were used with a diameter of 76.2 mm (3 in.), length of 910 mm (36

Page 96: RP 2002 Final Report - Georgia

83

in.) and wall thickness of 1.65 mm (0.065 in.). The crews from GDOT were told the foundation

depth of the bridges, and they chose the most convenient drilling method and the sampler

insertion method. Boring logs were provided for each of the sites where the samples were

extracted. After receiving the samples, they were sealed and stored in a constant temperature

room vertically confined inside a wooden box, until the soil and flume tests were ready to begin.

Ten bridge sites in the state of Georgia were chosen for collection of samples from their

foundations on which to perform flume tests and measure soil characteristics. Flume tests were

executed with the objective of finding the particular critical shear stress and erosion rate constant

at the respective site. A number of soil properties were measured including size distribution,

water content, bulk density, organic matter, and liquid and plastic limit for the fine-grained

samples. These sites were geographically distributed in such a way that they fell into different

river basins and into different physiographic regions.

Four main physiographic regions can be roughly identified in Georgia. They are the

Valley and Ridge, the Blue Ridge, the Piedmont and the Coastal Plain regions. Samples were

collected from each of these regions. Fig. 5.2 shows the location of the sites identified by the

county in which the bridge is located. Table 5-1 provides further information on the bridge

locations and physiographic regions in which they are found.

Page 97: RP 2002 Final Report - Georgia

84

Figure 5.2. Shelby tube core sample locations (Digital Environmental Atlas of Georgia, Alhadeff et al. 2000).

Table 5-1. Location of samples and description of physiographic regions.

Sample Number

County Location Physiographic Section

Major Land Resource Area

Latitude and

Longitude 1 Murray US 411 over Mill

Creek Southern Valley and Ridge Section

Southern Appalachian

34.8189º, 84.7647º

2 Towns SR 288 over Fodder Creek

Southern Blue Ridge Section

Blue Ridge 34.9275º, 83.7625º

3 Habersham Duncan Bridge Rd over Chattahoochee River

Southern Piedmont Section

Southern Piedmont 34.5406º, 83.6228º

4 Haralson US 27 over Tallapoosa River

Southern Piedmont Section

Southern Piedmont 33.8642º, 85.2097º

5 Wilkinson SR 57 over Oconee River

Sea Island and East Gulf Coastal Plain Section

Southern Coastal Plain

32.7817º, 82.9586º

6 Bibb US 80 / 5th St over Ocmulgee

Southern Piedmont Section

Sand Hill 32.8380º, 83.6212º

7 Effingham I-95 (NBL) over Savannah River

Sea Island Section Atlantic Coast Flatwoods

32.2351º, 81.1540º

8 Decatur SR 1b / Calhoun St over Flint River

East Gulf Coastal Plain Section

Southern Coastal Plain

30.9061º, 84.5886º

9 Berrien SR 76 over Withlacoochee River

East Gulf Coastal Plain Section

Southern Coastal Plain

31.1769º, 83.3225º

10 McIntosh US 17 over Darien River

Sea Island Section Atlantic Coast Flatwoods

31.3675º, 81.4364º

Page 98: RP 2002 Final Report - Georgia

85

In addition to the material collected from the ten sites, two samples were prepared to

calibrate the erosion measurements and to provide reference measurements on coarse sediments.

The first was bed material collected from Peachtree Creek inside the Atlanta metro area with low

clay content. The second reference material tested was a pure commercial sand having no silt or

clay content with a median size of 1.16 mm and a coefficient of uniformity of 1.5, which is the

ratio of the diameter of the particles corresponding to 60% and 10% finer on the cumulative

particle-size distribution curve.

Laboratory Flume Measurements

Shown in Fig. 5.3 is the rectangular, tilting, recirculating flume which was utilized to erode the

core samples. It is located in the hydraulics laboratory at the Georgia Institute of Technology.

The flow in the flume can be driven by either of two variable-speed pumps for low or high flows.

The flume is 20 ft long, 1.25 ft wide, and a maximum of 1.25 ft deep. The flume bed has fixed

small gravel (d50 = 3.3 mm) to assure a fully-developed boundary layer in fully-rough turbulent

flow. At the end of the flume there is a holding tank with a volume of 67 ft3, which feeds both

pumps. Only one pump is operated at a time. A 6 in. diameter pipe circulates the flow from the

large pump to the head box of the flume, which contains an elliptical wall transition and flow

stilling devices. On the other side of the flume, a 4 in. diameter pipe feeds water from the small

pump into the head box. The small pump is a progressing cavity pump for slurries while the large

pump is a low-speed, large-impeller centrifugal pump designed for solids pumping.

Page 99: RP 2002 Final Report - Georgia

86

Figure 5.3. Recirculating flume for erosion testing.

The operating variables of the flume are flow rate, slope and flow depth. The flow rate is

adjusted based on the rotational speed at which the pump is operating. Flow calibration tables

were developed in previous research (Ravisangar 2001 and Hoepner 2001). The 4 in. pump has

a working range from 0.12 – 0.52 cfs, while the 6 in. pump develops flow rates from 0.50 – 2.5

cfs.

Page 100: RP 2002 Final Report - Georgia

87

The tilting flume can be set at slopes between 0 and 0.02 ft/ft. The slope is measured by a

slope counter, which counts the revolutions of the gear mechanism that raises and lowers the

flume.

The flow depth is set using either the tailgate for subcritical cases or the upstream sluice

gate for supercritical flow. The values of the normal depth were determined in an initial set of

experiments as the asymptotic approach depth of gradually-varied flows. Depths were then set

to normal depth during erosion experiments because that guarantees a uniform flow and allows

calculation of the bed shear stress by the uniform flow formula. The resulting calculated shear

stresses were found to agree with values obtained from velocity profiles measured by a laser

Doppler anemometer as shown by Ravisangar (2001). The bed shear stresses ranged from 0.4 Pa

to 21 Pa (0.008 to 0.438 lb/ft2).

To measure the erosion rate of a sample, the Shelby tube is placed below a circular

opening in the bottom of the flume, and the position of the piston used to extrude the sample as

erosion occurs is tracked by a linear variable differential transformer (LVDT). As the material is

eroded over the exposed area of the Shelby tube sample, which is a circle approximately 3 in. in

diameter, the operator pushes the sample upward with the piston, maintaining the sediment

surface level with the top of the gravel bed of the rectangular flume. The material eroded is

recorded as the mass per unit area removed per unit time based on piston displacement as a

function of time measured by the LVDT and the measured dry density of the sample. Uniform

sand was tested in the flume, and the erosion rate values were found to be reproducible, while the

measured critical shear stress agreed with the Shields’ value within the range of experimental

uncertainty (Navarro 2004).

Page 101: RP 2002 Final Report - Georgia

88

RESULTS AND ANALYSIS

Observed Erosion Behavior

The observed erosion fell mainly in two of the three modes of erosion identified by

Mehta (1991). At low shear stresses, just above the critical shear stress, single particles were

dislodged over the entire bed, which is identified as surface erosion. The other type of erosion

observed in this study was mass erosion, which occurred at shear stress values greater than those

that created surface erosion. In this last case, the material failed along a plane, transporting all

the material above it. Even though these two mechanisms could be distinguished, there was not

a clear line of demarcation between them. These two mechanisms usually coexist but the

predominance of one over the other is likely to depend on the amount of fine material present in

the sediment and the size of the fine material.

Experimental Results

Among the many empirical relationships between erosion rate and shear stress for fine-grained

sediments that have been proposed, as summarized by Mehta (1991), three were explored in

detail: linear, exponential and power. Two of them, linear and exponential, showed the best

agreement with the experimental data and were included in the analysis based on their goodness

of fit and the standard error of the erosion parameters. For those relationships, seventeen

sediment samples out of the thirty-one samples on which soil classification tests were performed

had acceptable erosion rate vs. applied shear stress relationships. An acceptable relationship was

defined as one having a coefficient of determination greater than 0.50 (R2 > 0.50). This criterion

was applied to the two best regression models, which were piecewise linear and exponential

relationships.

Page 102: RP 2002 Final Report - Georgia

89

The linear model was first proposed by Kandiah (1974), and it is given by

−⋅=

c

cMEτ

ττ (5.3)

in which E = erosion rate; M = erosion rate constant; τ = applied shear stress; and τc = critical

shear stress. In this study, a piecewise linear relationship was developed to better fit the data. The

exponential model has been proposed in several forms (see Mehta 1991), but the form used in

this study is given by

−⋅

⋅= c

ca

c eEE τττ

(5.4)

in which Ec = critical erosion rate; a = erosion rate constant; τ = applied shear stress; and τc =

critical shear stress.

The critical shear stress is defined in the linear model by the extrapolation of the best-fit

line for erosion vs. applied shear stress to an erosion rate equal to zero. Flow conditions under

this critical value of shear stress produce insignificant erosion. In the exponential model, a value

of negligible erosion rate has to be specified in order to find the intercept and thus the critical

shear stress, given that it is an asymptotic model. The critical erosion rate for the exponential

model is defined in this study as the value of erosion rate that gives a minimum least squares

error between the critical shear stress values found by linear regression and by exponential

regression. This value of the erosion rate was found to be 0.00190 kg/m2/s, which is acceptable

given that it is approximately twice the value of the minimum erosion rate that could be

measured.

Fig. 5.4 shows the erosion test results for the first group of sediments for which measured

erosion rates reached 1.1 kg/m2/s. The solid lines are the best-fit linear and piecewise linear

models, and the dashed lines are the best-fit exponential models. Each sample is identified by

Page 103: RP 2002 Final Report - Georgia

90

the county in which it is located as given previously in Table 5-1 and by the depth of the sample

layer. Fig. 5.5 shows the measured erosion rates and their best-fit models for erosion rates up to

0.06 kg/m2/s. In some cases, the data follow a single linear relationship rather than a piecewise

linear one.

The values of critical shear stress and erosion rate constants for each sample are

summarized in Appendix A in Tables A-1 and A-2, respectively. The linear regression model

gives a standard error in the critical shear stress of 0.45 Pa compared to 0.57 Pa for the

exponential model, and so the critical shear stress from the linear model is used in subsequent

regression analyses. The relative uncertainty in the erosion rate constants for the two models is

29% for the linear model and 22% for the exponential model.

0.001900.0

0.2

0.4

0.6

0.8

1.0

1.2

0 5 10 15 20

Applied Shear Stress (Pa)

Eros

ion

Rat

e (K

g/m

2 /s)

Ptree Creek

Murray 3'-4'

Towns 7' - 8'

Habersham 12' - 12' 6"

Haralson 12' - 15' 5''

Effingham 21' - 22'

McIntosh 10' - 12'

Critical Erosion

Figure 5.4. Erosion rate vs. applied shear stress relationships for sediment samples with erosion rates up to 1.1 kg/m2/s (samples identified by county and depth except for Peachtree Creek in Atlanta metro area; see Fig. 5.2).

Page 104: RP 2002 Final Report - Georgia

91

0.001900.00

0.01

0.02

0.03

0.04

0.05

0.06

0 5 10 15 20

Applied Shear Stress (Pa)

Eros

ion

Rat

e (K

g/m

2 /s)

Towns 5' - 5' 6"

Towns 6' - 7'

Habersham 11' - 12'

Habersham 12' 6" - 14'

Wilkinson 36'6" -37'6"Bibb 25' - 25' 8"

Bibb 30' 8" - 31' 4"

Bibb 31' 4" - 32'

Decatur 20'-20'4"

Critical Erosion

Figure 5.5. Erosion rate vs. applied shear stress relationships for sediment samples with erosion rates up to 0.06 kg/m2/s (samples identified by county and depth; see Fig. 5.2).

Overall, the linear regression model, or piecewise linear regression model, showed good

agreement with the data. This type of model is certainly easier to manipulate and fewer

parameters have to be defined, which makes it more robust to apply in predictions. However, the

range of applicability of the linear models is more limited. For larger ranges of shear stress

values, piecewise linear models are required, although this increases the number of parameters to

specify. The location of the break point for the piecewise linear model is based on judgment

applied to the plotted data and on maximizing the coefficients of determination of each segment.

Linear models are preferred for low increments of the applied shear stresses beyond the critical

value, and they are more accurate in finding the critical shear stress values using the first part of

the piecewise relationships.

Page 105: RP 2002 Final Report - Georgia

92

The exponential model describes a more realistic shape of the erosion rate function over

the full range of low to high shear stresses. This model is asymptotic to the shear stress axis and

an additional variable, the critical erosion rate, has to be defined. The critical erosion rate value

acts as an axis intercept for the model in order to find the critical shear stress. The value of the

critical erosion rate has to satisfy the physical restriction of the minimum laboratory erosion rate

that can be measured in the laboratory apparatus, as well as provide a consistent intercept for

determining the critical shear stress in the exponential model.

Results for all of the sediment properties measured for each sample are also given in

Appendix A in Tables A-3 through A-12 along with a map of the sample location, a

classification into soil types, and the relative percentage by weight of sand, silt, and clay in each

sample. The sediment properties included in the tables are dry density, void ratio, bulk density,

water content, specific gravity, organic matter, liquid limit, plasticity index, and median grain

size.

Prediction of Critical Shear Stress and Erosion Rate Constant

The best group of independent variables for prediction of critical shear stress and the erosion rate

constant of the sediment samples was determined using MINITAB statistical software to perform

simple and multiple linear regression analysis.

Among the sediment parameters measured for all the sites, the following were included in

the statistical analysis: bulk density (kg/m3), water content (decimal fraction), organic matter

content (decimal fraction), median sediment size (mm), clay content (decimal fraction) and fines

content, defined as the sum of the clay and silt content (decimal fraction).

The soil parameters that were measured but not included in the statistical analysis were

the specific gravity, the liquid limit, and plastic limit. The specific gravity was excluded because

Page 106: RP 2002 Final Report - Georgia

93

of its low variability. The liquid limit and plasticity showed some correlation since, for example,

the most resistant material (Berrien Co.) had by far the largest values of liquid and plastic limit.

However, only eight of the seventeen data points were classified as plastic material, too few to

perform confident regression analysis.

In order to verify the goodness of fit of the regression models, statistics such as the

coefficient of determination, the adjusted coefficient of determination, the Mallow’s Cp, and the

estimated standard error of the model were evaluated.

Critical Shear Stress

The statistics mentioned above are utilized to assess the goodness of fit and the goodness of

prediction of a regression model in order to choose the best model for prediction of critical shear

stress. In the first multiple linear regression model considered, the predictors selected are

• bulk density (kg/m3),

• water content (decimal fraction),

• organic matter (decimal fraction),

• median size (mm),

• clay content (decimal fraction), and

• fines content (decimal fraction).

The response variable is the critical shear stress found from the intercept of the first segment of

the piecewise linear regression relationship between erosion rate and shear stress.

For the linear model applied to critical shear stress, one outlier point was identified,

which was the sample from McIntosh County. The material from this site can be considered

unusual because of its content of shells. After removing this point, fines content is the best

predictor, followed by organic matter and median sediment size. It should be mentioned that the

Page 107: RP 2002 Final Report - Georgia

94

clay content was expected to be a better predictor variable than fines content given that at clay

sizes the change to platelet-like shape magnifies the interparticle forces. Also, Kamphuis and

Hall (1983), who performed initiation of motion tests on consolidated fine sediments, found that

the shear stress required to initiate motion increases with increases in the clay content. However,

for the data obtained in this study, the fines content is a better predictor variable.

Given the limited number of data points (16), it was decided to use a maximum of three

predictors in a regression model. Considering the values of the model statistics, the models with

three variables perform better, and the best-fit linear model with three predictors is given by

5037.24.617.2576.0 dOMFinesc ⋅+⋅−⋅+=τ (5.5)

in which τc = critical shear stress, Pa; Fines = decimal fraction of fine material by weight; OM =

decimal fraction of organic matter by weight; and d50 = median grain size, mm.

The performance of the regression equation of measured vs. predicted shear stress gives a

value of R2 = 0.72, and the standard error of estimate in the critical shear stress is 3.1 Pa which is

greater than the estimated experimental uncertainty of 0.5 Pa. This means that there is additional

unexplained variation that cannot be accounted for by the experimental uncertainty.

An alternative three-parameter model with about the same performance as Eq. 5.5

brought in bulk density as a replacement for median grain size. However, the regression equation

shows a decrease in critical shear stress with an increase in the bulk density, which contradicts

the results obtained by other researchers (Mehta 1991, Krone 1999, Ravisangar et al. 2001, and

Briaud et al. 2001). They have found that a more compact or denser fine material will better

resist erosion. This contradiction can be explained for the sediments tested in this study, which

are a mixture of fine and coarse sizes, by the different bulk density values of the sand and the

clay. Consider that the bulk density of pure sand is higher than for clay. In the case of pure clay

Page 108: RP 2002 Final Report - Georgia

95

samples, they become more resistant as their density increases as found by other investigators.

However, for mixtures of clay and sand material having higher bulk density than clay alone, the

critical shear stress may not be higher because of the presence of sand which reduces the

interparticle forces. For this reason, bulk density is not a clear predictor variable for mixed

sediment layers.

A third regression model that was tried was a two-variable model that utilized

nondimensional variables for comparison with the Shields diagram for coarse sediments in the

form given by Fig. 5.1 with τ*c = Shields parameter = τc/(γs − γ)d50 as a function of the

dimensionless particle diameter d* = [(SG − 1)gd503/ν 2]1/3 where SG = specific gravity of the

sediment and ν = kinematic viscosity of the fluid.

Unlike Shields’ data, the natural sediment exposed to erosion around bridge foundations

is a mixture of both fine and coarse sediments with varying magnitudes of interparticle forces

that can affect the comparison. The best two-variable predictor model that includes log d* also

includes Fines content (decimal fraction) as the second variable. This analysis includes all 17

data points and explains the behavior of the sample from McIntosh Co. The regression equation

is given by

337.0*

67.2* 10586.0 −⋅ ⋅⋅= dFinescτ (5.6)

For this regression relationship, the standard error in the log of the Shields parameter is 0.3 and

R2 = 0.89. It is of interest to note that experimental data for silt-size particles (crushed quartz)

plots on the Shields diagram with an exponent on d* of −0.39 which is close to the value in Eq.

5.6 (Sturm 2001). Fig. 5.6 shows the comparison between the measured and predicted critical

Shields parameter using Eq. 5.6.

Page 109: RP 2002 Final Report - Georgia

96

0.01

0.1

1

10

100

0.01 0.1 1 10 100

Predicted Shields Parameter, τ *c

Mea

sure

d Sh

ield

s Par

amet

er, τ

*c

R2 = 0.89

Figure 5.6. Comparison of the measured and predicted critical shear stress parameter using Eq. 5.6.

Fig. 5.7 shows the results plotted using the Shields diagram coordinates given previously

as Fig. 5.1. It can be observed that the measured values close to the Shields curve correspond to

sandy material with low fines content. For the special case used to calibrate the flume testing

with uniform sand material, which is similar to the material used to develop the Shields curve,

the data point falls within the upper range of the Shields curve. Either Eq. 5.6 or Fig. 5.7 is

recommended to estimate the critical shear stress value of sediment when its fines content and

size are known.

Page 110: RP 2002 Final Report - Georgia

97

10% fines

30%

50%

70%

90%

0%

74%

56%52%50%

56%40%

17%29%22%

25%5%

7%1% 3%

13%10%

0.01

0.1

1

10

100

1000

0.01 0.1 1 10 100 1000Dimensionless Diameter, d * = [(SG-1)gd 50

3 / ν 2 ] 1/3

Shie

lds P

aram

eter

, τ*=

τ c/(γ

s − γ

)d50

Shields Measured

Figure 5.7. Comparison of measured data and calculated values using Eq. 5.6 plotted on Shields’ diagram format.

Erosion Rate Constant

For excess shear stress relationships, the second parameter of importance is the erosion rate

constant. This constant, called “M” in the linear model (Eq. 5.3), and “a” in the exponential

model (Eq. 5.4), quantifies the relative increase in the erosion rate as the response for an increase

in the applied shear stress above its critical value. In the linear case, M is defined as the erosion

rate predicted for an applied shear stress equal to twice the value of the critical shear stress.

Because of the relatively poor correlation found between the erosion rate constants a or M and

sediment properties, a possible correlation was also sought using slightly different expressions as

suggested by Lee et al. (1994) for the erosion rate dependence on shear stress. They proposed a

linear model of the form

)(1 csE ττ −= (5.7)

Page 111: RP 2002 Final Report - Georgia

98

in which s1 represents the erosion rate predicted for a unit increment above the critical shear

stress. Similar definitions can be stated for the exponential case. The proposed erosion rate

relationship for the exponential case becomes

)(2 csceEE ττ −= (5.8)

Six possible forms of the erosion rate constants given by M, s1, log s1, a, s2, and log s2 were

studied. The best expression found for the erosion-rate constant in the linear case is using the

logarithm of the response variable, log s1, where s1 is in (kg/m2/s)/Pa. The best predictors are the

logarithm of the fines content, log Fines, where Fines is given as a decimal fraction, and the

dimensionless particle diameter, d*. The expression is given by

*0305.011.11 1000191.0 dFiness ⋅− ⋅⋅= (5.9)

which applies for values of Fines > 0. Eq. 5.9 has a coefficient of determination of R2 = 0.79, and

a standard error in log s1 of 0.42.

In Fig. 5.8, Eq. 5.9 for the erosion rate constant s1 is compared with the laboratory data

using a plot similar to the Shields diagram. The dimensionless diameter d* is plotted on the x-

axis, while on the y-axis the erosion constant s1 replaces the value of the dimensionless shear

stress. It is of interest to note that the values for s1 found by Hoepner (2001) for an estuary mud

collected from the Providence River in Rhode Island agree well with the proposed relationship.

This material had values of d* between 0.25 and 0.5, and Fines content between 84 and 99%.

Page 112: RP 2002 Final Report - Georgia

99

100%

32%

10%

3%

56%

50%52%

74% 56%40%

17%

7%10%

22%

5%

29%

25%

1%

13%

3%

1% fines

estuary mud

0.001

0.01

0.1

1

10

0.1 1 10 100Dimensionless Diameter, d * = [(SG-1)gd 50

3 / ν 2 ] 1/3

Eros

ion

Con

stan

t, s 1

, (kg

/m2 /s

)/Pa.

Measured estuary mud (Hoepner 2001)

Figure 5.8. Comparison of the measured data and calculated values for erosion rate constant s1 using Eq. 5.9.

Analyzing the results in Fig. 5.8, it can be observed that for d* < 1.5, which corresponds

to the upper limit of the silt-size range, the erosion rate constant depends only on the fines

content. For larger values of d*, it depends on both fines content and d*.

The best-fit relationship found to estimate the erosion rate constant s2 for the exponential

model is linear. It relates the value of s2 with Fines and d*, and it is given by

*2 0794.060.144.1 dFiness ⋅+⋅−= (5.10)

which applies for materials with Fines content higher than 5%. Four out of the thirteen

measured relationships between erosion rate and shear stress had a poor exponential fit. The

coefficient of determination for the best-fit relationship given by Eq. 5.10 is R2 = 0.74. Although

Fines content does not seem to play a very important role according to the statistical output, it is

included as a predictor given that it has been shown to be the most important variable in the

Page 113: RP 2002 Final Report - Georgia

100

analysis of both the critical shear stress and the erosion rate constant for the linear model. The

results are presented in graphical form in Fig. 5.9. The regression relationship seems to follow

the same trends as for s1 shown previously in Fig. 5.8.

50%

90%

50%56% 52%

74% 56%17%

40%29%

10%

22%

25%

13%

3%

10% fines

0

0.5

1

1.5

2

2.5

3

3.5

4

4.5

0.1 1 10 100Dimensionless Diameter, d * = [(SG-1)gd 50

3 / ν 2 ] 1/3

Eros

ion

Con

stan

t, s 2

, Pa

-1.

Measured

Figure 5.9. Comparison of the measured data and calculated values for erosion rate constant s2 using Eq. 5.10.

Discussion

Given the limited experiments performed, it must be emphasized that these results are applicable

only to sediments collected from similar environments. The material studied came from bridge

foundations at depths from 1 to 35 ft deep below the top of the river bed and was usually

consolidated. The relationships that were developed to estimate the critical shear stress and

linear erosion rate constant are encouraging. The relationship found for the exponential erosion

rate constant is less reliable than the others, but it has the advantage of describing the erosion rate

over a wider range.

Page 114: RP 2002 Final Report - Georgia

101

In studying the geographic distribution of the results for critical shear stress given in

Appendix A along with the sediment properties, it is difficult to discern definitive regional trends

in the data. What is clear, however, is the considerable stratification with depth of erosion and

other sediment properties for the same sample. In general, the critical shear stress in the Valley

and Ridge and the Piedmont Provinces ranges between approximately 3 and 10 Pa (0.063 and

0.21 lbs/ft2) as long as the percent fines is less than about 50%, but considerable variability is

obvious within each sample and from sample to sample. On the other hand, in the Sea Island

Section and the East Gulf Coastal Plain, larger percentages of clay can give rise to very high

values of critical shear stress. As shown in this chapter, the unifying explanatory variables for

critical shear stress are the median sediment size and the percent fines, which can be used along

with the results presented in Appendix A to provide general guidance on scour resistance. In the

case of high-risk bridges, individual site-specific erosion tests are recommended.

Page 115: RP 2002 Final Report - Georgia

102

CHAPTER 6. COMPARISONS OF NUMERICAL MODEL, LABORATORY, AND FIELD RESULTS

INTRODUCTION

First, in this chapter, the numerical model is validated by presenting comparisons between

measured and computed mean velocity and turbulence kinetic energy profiles. Due to space

considerations, detailed comparisons are shown only for the Chattahoochee River site near

Cornelia and only a small sample of comparisons is included for the Ocmulgee and Flint River

sites. It should be noted, however, that for all three cases the comparisons between predictions

and laboratory measurements lead to essentially the same conclusion; namely, that the numerical

model can capture most experimental trends with good accuracy. After presenting the validation

of the numerical model the computed flowfields are analyzed in tandem with the laboratory

scour experiments to elucidate the physical mechanisms that contribute to local scour and to help

interpret the laboratory results.

Secondly, laboratory and field results are compared for a measured bank-full flow at the

Chattahoochee River site and an extreme historical flood event at the Flint River site (the 200-yr

flood resulting from Tropical Storm Alberto in 1994). These comparisons focus on the river-

model results rather than the flat-bed model results because the full three-dimensional

bathymetry is reproduced in the river models. In the course of these comparisons, the laboratory

scour and velocity measurements are validated with the field data and further insights into the

laboratory scale-up problem are discussed.

Page 116: RP 2002 Final Report - Georgia

103

NUMERICAL MODEL VALIDATION

Chattahoochee River near Cornelia

The flow around a single bent of the Cornelia site bridge piers mounted on a flat bed are

extensively studied both numerically and experimentally. In the experiments, the mean velocity

and turbulence kinetic energy distributions for the flat-bed model near the bridge foundations

were measured in addition to the scour hole bathymetry at equilibrium. Before presenting the

comparisons, it is important to mention that a series of numerical sensitivity studies were carried

out in order to demonstrate that the numerical results are free of numerical errors due to

insufficient grid resolution and other numerical parameters. For that reason, calculations were

carried out on three different grid systems with different grid resolutions and different sizes of

flow domains. Comparisons of the results from these different grid systems show that all of them

agree well with each other and approach the same result. Therefore, in the following

comparisons, numerical solutions are given for only one grid system.

Fig. 6.1 shows the comparison of measured and computed streamwise velocity profiles at

various locations upstream, within, and downstream of the piers (for detailed comparison

locations see Fig. 4.5(c)). The profiles show velocity variations in the transverse, y, direction at

various streamwise locations and at three different depths. It is important to note that the

measured velocity profiles are not perfectly symmetric and uniform. Moreover, the degree of

non-uniformity in the measured profiles appears to vary with depth. On further investigation of

the measured direction of the approach velocity vectors relative to the piers, it was discovered

that there was a slight skewness of about 1.8° clockwise in the approach velocity vector relative

to the pier centerline. This may have been due to inherent construction tolerances in setting the

piers relative to the approach flow or due to a slight asymmetry in the approach flow itself. This

Page 117: RP 2002 Final Report - Georgia

104

small degree of incoming flow skewness is hardly detectable in flow visualization. However, in

order to examine its effects on the flow, especially the effects on turbulence structures, flows

with different incoming directions were investigated. In the first case, the flow is aligned with

the bridge pier centerline and in the second case, the flow is skewed 1.8° clockwise.

Comparisons of the simulated and measured velocity profiles are shown in Fig. 6.1 in which the

red line depicts the aligned case while the green one corresponds to the skewed flow simulation.

Fig. 6.1 reveals that the numerical model captures most trends observed in the experiments with

very good accuracy. For instance, both the reduction of the centerline velocity as the flow

approaches the pier bent and the growth of the wake between and downstream of the piers are

predicted with good accuracy by the numerical model. It is also observed in the comparisons in

Fig. 6.1 that effects of the flow skewness can only be seen in the more downstream portion of the

bridge pier bent, with larger discrepancies in the wake region.

Figure 6.1. Comparisons of streamwise velocity profiles (location F1, F2, …, F6 left to right) at relative elevations of 0.6, 0.4, and 0.2 times the depth from top row downward (circles: lab data; red curves: aligned flow; green curves: skewed flow).

Page 118: RP 2002 Final Report - Georgia

105

In stark contrast to the relatively small effect of the flow alignment on the mean velocity

distribution, the experimental and computational results reveal a rather dramatic effect of even a

small misalignment on the turbulence structure in the vicinity of the piers. Fig. 6.2 shows the

calculated and measured profiles of total turbulence kinetic energy in the vertical direction at

several streamwise locations to the left and right of the bridge pier bent (locations are marked as

P1 to P4 in Fig. 4.5(c)). A remarkable feature of the measured kinetic energy profiles, which are

shown by symbols, is the large asymmetry of the turbulence structure with respect to the

streamwise axis of symmetry of the pier bent. At the first location (P1) the measured kinetic

energy profiles to the right and left of the pier bent are nearly identical and in good agreement

with the simulations. Further downstream, however, the kinetic energy to the right side of the

pier bent rises sharply in the outer layer yielding a highly asymmetric turbulence structure. The

numerical simulations of the aligned flow case, on the other hand, yield a symmetric turbulence

structure, which is to be expected since the simulated conditions are perfectly symmetric.

However, unlike the limited effects on the time-averaged velocity distribution, the slight

misalignment of incoming flow has a considerable effect on the distribution of the turbulence

kinetic energy. As depicted by the green lines, the simulated result of the skewed flow exhibits

qualitatively the main trend observed in the data: the steep rise of turbulence kinetic energy on

the right side of the piers and the gross asymmetry of the turbulence structure.

Page 119: RP 2002 Final Report - Georgia

106

Figure 6.2. Comparisons of turbulence kinetic energy (circles: lab data, red curves: aligned flow; green curves: skewed flow).

As suggested by the above results, the numerical model can capture the apparent

sensitivity of the turbulence structure to even a relatively small misalignment of the pier bent

with respect to the approach flow. Yet, significant discrepancies between the measured and

computed turbulence kinetic energy profiles remain even for the skewed flow simulation. These

discrepancies may be due to the fact that the exact state of the approach flow in the laboratory

experiment is not known. As indicated by the mean velocity profiles shown in Fig. 6.1 at the

most upstream measurement stations, the approach flow in the laboratory exbibits small but

clearly visible spanwise variations. Unless these variations can be quantified in sufficient detail

to be incorporated into the model boundary conditions it is unlikely that improved predictions of

the turbulence structure can be obtained.

Flint River at Bainbridge

The flow around a single bent of the Flint River bridge piers mounted on a flat river bed is also

studied both numerically and experimentally. In the experiments, the mean velocity distribution

and equilibrium scour depths were measured for the flat bed near the bridge piers. The first

simulations shown here are based on the initial state of the river bed in which it is flat and the

Page 120: RP 2002 Final Report - Georgia

107

footings are not exposed. The following comparisons are focused on the numerical solution with

a grid discretization consisting of 4 subdomains containing approximately 1×106 elements. Fig.

6.3(a) shows the two grids around the piers and the intermediate grid connecting them. The

locations of the six cross-sections where results of the simulations are compared with

experimental data are shown in Fig. 6.3(b).

Fig. 6.4 shows the comparison of measured and computed streamwise velocity profiles at

six locations upstream, within, and downstream of the piers (sections S1 to S6). These

comparisons are similar to those for the Chattahoochee River site. The comparison between the

profiles of measured and computed velocities, at three vertical positions near the bed, shows

reasonably good agreement. The largest disparities between the computed and observed profiles

occur at section S4 between the piers at 30% of the total depth, but the wake comparisons at

section S6 are quite good.

Figure 6.3. Flint River Bridge Layout (a), and measurement cross-sections (b).

Page 121: RP 2002 Final Report - Georgia

108

Overall, for the Flint River Bridge, the numerical model can capture the characteristics of

the mean velocity profiles observed in the experiments, especially the reduction in the velocity at

the centerline and the velocity profiles in the wake of the piers.

At the equilibrium scour state, the bridge footing is exposed to the flow. This change may

drastically change the flow status and the capacity for scouring; therefore, it is of fundamental

importance to study the flow with footings exposed. A new case, in which the footings along

with the bridge piers are mounted on a flat river bed, is generated to investigate the more

complex flow. As shown in the instantaneous streamlines for this simulation, depicted in Fig.

Figure 6.4. Comparisons of streamwise velocity profiles at locations from S1 to S6

shown in Fig. 6.3 (left to right) and at water depths of 0.45, 0.3, and 0.1 times the

depth from top to bottom row (circles: measurements; curves: numerical simulation).

Page 122: RP 2002 Final Report - Georgia

109

6.5, the foundations of each pier have a considerable influence over the flow patterns at the bed,

producing recirculating zones behind the footings and decreasing the velocity in the wake. It is

worthwhile to note that the river bed is not flat at equillibrium scour, and the actual flow may

differ from what is presented here. Nevertheless, this simulation once again illustrates the strong

capability of the 3D numerical model to handle complex pier geometry.

Ocmulgee River at Macon

A single bent of four cylindrical piers at the Ocmulgee River bridge in Macon GA is simulated

over a flat bed, in which the solution domain is discretized in 5 overlapped subdomains as seen

in Fig. 6.6. The complexity of the flow features in this simulation required an increase in the grid

resolution in order to find a solution free of numerical errors. Satisfactory results were obtained

with a grid consisting of approximately 2×106 elements.

Figure 6.5. Snapshot of the streamlines at the Flint River bridge,

including the footings in the simulation.

Page 123: RP 2002 Final Report - Georgia

110

From the results it can be observed that the flow features obtained in this simulation are

similar to the other two cases. The flow is characterized by unsteady coherent structures which

form a shear layer in the direction parallel to the pier bent. In Fig. 6.7(a), instantaneous

streamwise velocity contours show periodic vortex shedding in the wake flow. Fig. 6.7(b) shows

the typical unsteady tornado-like vortices, which appear and disappear continuously behind the

piers. Time-averaged streamwise velocity profiles are shown in Fig. 6.8.

(a) Streamwise velocity contours. (b) 3D streamlines.

Figure 6.7. Visualization of instantaneous flow field in the vicinity of the bridge piers.

(a) (b)

Figure 6.6. Geometry of the four overlapped grids around a single bent

of the Ocmulgee River bridge piers .

Page 124: RP 2002 Final Report - Georgia

111

Figure 6.8. Time-averaged streamwise velocity profiles at locations S1, S2, and S3 shown in Fig. 6.6 (left to right) and at water depths of 0.5, 0.3, and 0.1 times the depth from top to bottom row.

Experimental measurements of velocity profiles for the Ocmulgee River site will be

conducted in the GDOT Phase 2 project, and detailed comparisons between experiments and

numerical simulations will be made.

FLOW STRUCTURES AND SCOUR In this section, links are established between the complex hydrodynamics induced by the bridge

piers, as obtained from the numerical simulations, and the scour patterns that result under the

same flow conditions in a laboratory experiment with the same piers installed on an erodible bed.

Since the numerical computations have assumed a fixed flat bed, the discussion herein is only

qualitative. It is aimed at underscoring the complexity of the hydrodynamic processes that drive

the scouring process in real-life bridge foundations and at providing guidance for future

extensions of the model to develop a numerical scour-prediction tool. All the subsequent

discussion will focus on the Cornelia site.

Page 125: RP 2002 Final Report - Georgia

112

The equilibrium scour patterns obtained from the laboratory experiments are shown in

Fig. 6.9. As seen in this figure, a scour trench develops that surrounds the entire foundation with

the deepest scour occurring upstream of the first pier. Another region of relatively deep scour

within this trench is also observed just upstream of the last pier. It is important to note the overall

asymmetry of the scour patterns, which becomes more pronounced downstream of the first pier.

Such asymmetry is in accordance with the previously discussed impact of approach flow

skewness on the structure of the foundation-induced turbulence.

Most available sediment transport models employ the concept of critical bed shear stress,

the so-called Shields parameter, to define the threshold for incipient sediment grain motion. It

would, thus, be instructive to examine the simulated bed shear stress contours for the flat-bed

case because that would tend to identify regions in the flow where the scouring process is

initiated. The calculated time-averaged shear velocity contours are shown in Fig. 6.10(a). Two

pockets of maximum shear velocity are observed at the two upstream corners of the first pier.

The calculated shear velocity levels within these pockets are at least one order of magnitude

greater than the shear velocity levels within the rest of the foundation. This trend is to be

expected since the last three piers are embedded within the wake of the first pier and the flow in

their vicinity is, thus, dominated by large-scale, three-dimensional separation and flow reversal.

The pockets of large shear velocity correlate well with the region of maximum scour depth

surrounding the first pier.It is evident from Figs. 6.9 and 6.10(a), however, that the distribution

of bed shear stress alone cannot account for the complexity of the scour patterns observed in the

experiment. To further elucidate the role of foundation-induced hydrodynamics on scour, Fig.

6.10(b) shows contours of vertical time-averaged velocity at a horizontal plane very close to the

channel bottom (0.01H). The vertical velocity component is a good indicator of the complexity

Page 126: RP 2002 Final Report - Georgia

113

Figure 6.9 Distribution of scour depth at the equilibrium state.

(a) Friction velocity. (b) Vertical velocity.

(c) Limiting streamlines.

Figure 6.10 Flow patterns near river bed.

Page 127: RP 2002 Final Report - Georgia

114

and three-dimensional structure of the vortical patterns near the foundation. For example, a

pocket of negative vertical velocity component near a pier indicates that the flow along the

obstacle is directed toward the bed. For continuity to be satisfied, however, such a pocket of

downflow must be accompanied by a horizontal flow along the bed directed away from the

obstacle, which would tend to sweep bed material away from the obstacle and promote scour.

Alternatively, a pocket of positive vertical velocity around a pier suggests a vertical upwelling

along the pier away from the bed and must be accompanied by a region of horizontal flow

directed toward the obstacle. Such secondary flow patterns would tend to sweep bed material

toward the obstacle and lead to local deposition. To better illustrate these flow patterns at the

horizontal plane, Fig. 6.10(c) shows the limiting streamlines (or skin-friction lines)

corresponding to the vertical velocity contours shown in Fig. 6.10(b). As seen in Figs. 6.10(b)

and (c), the region of negative vertical velocity around the first pier is indeed accompanied by a

horizontal flow along the bed directed away from the pier. The topology of the limiting

streamlines in this region, which consists of the C-shaped separation line surrounding the

obstacle, the saddle node delineating the approach and near-obstacle flows, and the half saddle

node on the upstream face of the obstacle, is characteristic of the horseshoe vortex system

induced by the pier. Similarly, the pocket of positive vertical velocity at the downstream end of

the first pier is indeed accompanied by a near-wall flow directed toward the pier, which emanates

from the half saddle node on the upstream face of pier number 2. It is also worth noting from

Fig. 6.10(c) the complexity of the topology of the limiting streamlines around piers 2, 3, and 4,

which is characterized by the presence of pairs of saddle foci in the wake of each pier. These

saddle foci tend to sweep flow toward each pier and are thus the footprints on the bed of

vertically oriented, tornado-like vortices.

Page 128: RP 2002 Final Report - Georgia

115

Juxtaposing now the near-bed flow patterns shown in Fig. 6.10(a)-(c) with the observed

scour map shown in Fig. 6.9 reveals that the region of deepest scour at the front of the upstream

pier (pier 1) correlates well with the pocket of negative vertical velocity, the associated region of

near-bed flow away from the pier, and the two pockets of maximum bed shear stress. The second

region of deep scour located at the face of the most downstream pier (pier 4) correlates well with

the pocket of negative vertical velocity even though no appreciable levels of shear velocity exist

in this region.

Another interesting feature of the scour patterns visible in Fig. 6.9 is the characteristic C-

shaped structure of the bed-elevation contours at the downstream end of piers 1 and 4, which

reveals the presence of two small ridges of local sediment deposition with less scour adjacent to

the ridges. These ridges appear to correlate well with the pockets of positive vertical velocity in

the downstream end of piers 1 and 4, thus, supporting the previous qualitative discussion on the

role of local hydrodynamics in the sediment transport processes.

As remarked at the start of this section, the discussion herein is only qualitative. The

simulated flow patterns for the flat bed case can only provide some indication as to where and

how scour will originate. The complex deformation of the channel bed in the vicinity of the

foundation, as revealed by the experiments, will undoubtedly alter the local hydrodynamics

which will in turn affect the rate of sediment transport and deposition. The discussion in this

section, however, serves to clearly underscore that simplistic sediment transport models relying

exclusively on the concept of shear stress in excess of critical bed shear stress may not be

adequate for modeling scour at real-life bridge foundations.

Page 129: RP 2002 Final Report - Georgia

116

COMPARISONS OF LABORATORY AND FIELD RESULTS

Chattahoochee River near Cornelia

The comparisons in this section rely on a modeling strategy that is the outcome of the

comparisons made between laboratory scour measurements and accepted pier scour formulas in

Chapter 3. The 1:40 scale laboratory river model was constructed as a Froude-number model

with equality of y1/b values. The sediment size was selected to be 1.1 mm to obtain clear-water

scour near the maximum of V1/Vc = 1.0 at approximately the same Froude number as the

prototype for the bank-full and the 100-yr flood flows. Approach Froude numbers, Fr, do not

change very much for this range of events. The model sediment size results in a value of b/d50 =

24.5 in the laboratory at which several pier scour formulas indicate almost no effect of this

parameter. The comparisons discussed next refer to experimental run RM 5 in Tables 3-1 and 3-2

for which y1/b = 4.0, b/d50 = 24.5, V1/Vc = 0.75, and Fr = 0.30. The corresponding values of the

dimensionless parameters in the field for the bank-full event of July 2, 2003 are y1/b = 4.0, b/d50

= 1570, V1/Vc = 4.38, and Fr = 0.33.

Laboratory measurements of scour contours and velocity vectors (before scour) at a

relative height above the bed of 0.4 were shown previously in Fig. 3.10 for the bank-full flow of

13,600 cfs that occurred on July 2, 2003. The flood recurrence interval for this event is

approximately 2 years. The field fathometer measurements of bed elevation with time were

shown previously in Fig. 2.3 throughout the hydrograph for this event. The greatest scour occurs

in front of the nose of the upstream pier in agreement with the laboratory results, but there is an

obvious infilling of the scour hole on the recession side of the hydrograph after a constant

elevation is reached indicating equilibrium live-bed scour. Relatively little scour occurs on the

Page 130: RP 2002 Final Report - Georgia

117

right side of the upstream pier, but measurable scour is apparent around the sides of the most

downstream pier.

Measured channel cross sections just upstream of the bridge for several flow events are

compared with the laboratory experiments in Fig. 6.11. The event of June 13, 2003 was a very

small one, but it established the reference bed elevation of 1126 ft prior to the occurrence of the

flood event on July 2, 2003. There is relatively close agreement between the field cross sections

for the events of 1961 and July 2, 2003 which had almost identical discharges. Good agreement

is also shown in Fig. 6.11 between the laboratory cross section measured after scour and the field

cross sections for these two flood events measured near the time of peak discharge. The obvious

disagreement is the occurrence of what are apparently dunes to the left of the pier for the live-

bed scour in the prototype because the laboratory model measurement was taken for clear-water

scour.

1105

1115

1125

1135

50100150200250300

Station, ft

Ele

vatio

n, ft

Experiment(Q=1.35cfs)12/12/1961(Q=13100cfs)7/2/2003(Q=13600cfs)6/13/2003

EL 1126 ft

Right BankLeft Bank

Figure 6.11 Comparison of scour in prototype and laboratory cross sections at Cornelia site just upstream of the bridge looking downstream (field events and model Run RM 5).

Page 131: RP 2002 Final Report - Georgia

118

Comparisons between laboratory and field measurements of velocity to the left of the

central bridge pier are shown in Fig. 6.12 for the bank-full event of July 2, 2003. There is close

agreement between the laboratory velocities scaled up with Froude number similarity and the

field measurements made during the flood with the fixed acoustic Doppler instrument.

0

2

4

6

8

13.9 ft23.8ft33.6ft

Distance from left side of pier

Vel

ocity

, ft/s

Field measurement (7/2/2003) River model (Q=1.35 cfs)

2.3% 1.3%% diff.= 0.8%

Figure 6.12. Comparison of velocities between field and laboratory measurement at given distances from left side of the central pier bent.

The dimensionless maximum pier scour depths are shown in Fig. 6.13 for three

laboratory clear-water scour experiments with differing values of V1/Vc all less than 1.0 but with

constant values of y1/b = 4.0 in agreement with the bank-full flood event. The values of the

Froude number are shown next to each data point. The data point shown in Fig. 6.13 with a

laboratory Froude number of 0.30 is the one that represents the laboratory results that have been

compared favorably with field data in all previous figures, and it agrees relatively closely with

the prototype Froude number of 0.33. Also shown in Fig. 6.13 are the pier scour formulas of

Melville (1997) and Sheppard (2003) for clear-water scour in the laboratory with b/d50 = 24.5

Page 132: RP 2002 Final Report - Georgia

119

and y1/b = 4.0. There is good agreement between the laboratory data and these two clear-water

scour formulas. However, the prototype is in the live-bed scour regime for the bank-full event

with d50 = 0.7 mm and b/d50 = 1570, which is obviously quite different than the model value.

Accordingly, the proposed live-bed scour formula of Sheppard (2003) obtained from scour data

in a large flume is compared with the field scour depth in Fig. 6.13, and the results agree

reasonably well considering that Sheppard’s formula has been extrapolated beyond the

maximum range of his data of b/d50 = 564. His large-flume data suggest that very large values of

b/d50 which occur in the field diminish the maximum clear-water scour depth with approximately

a straight line drawn by his formula between the reduced maximum clear-water scour depth and

the live-bed scour peak at which the bedforms become flat or plane bed.

0.0

1.0

2.0

3.0

4.0

5.0

0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0V 1 /V c

d s/ b

Lab River model, b/d50=24.5Sheppard (2003), b/d50=24.5Melville (1997) , b/d50=24.5Field Measurement, b/d50=1570Sheppard (2003), b/d50=1570

0.30Fr=0.33

0.400.33

Figure 6.13. Comparison of field and laboratory measurement of scour depths and Sheppard’s equations for Chattahoochee River, near Cornelia, GA (y1/b=4.0).

Page 133: RP 2002 Final Report - Georgia

120

Flint River at Bainbridge

A modeling strategy similar to that adopted for the Chattahoochee River site was applied

to the Flint River site for the historical event associated with Tropical Storm Alberto in 1994

which was a 200-yr event (107,000 cfs). In contrast to the bank-full event modeled for the

Chattahoochee River, this site allowed the modeling of an extreme flood event for comparison

with historical field measurements. The sediment size was 1.1 mm in a 1:90 scale model with a

Froude number similar to that in the field but in the clear-water scour regime. The 1:90 scale was

necessary to capture the very wide floodplains at this site. In addition, the upstream railroad

bridge and the abandoned embankment of the old highway bridge were included in the model in

order to reproduce the approach flow conditions as closely as possible (see Chapter 3). The

comparisons discussed next refer to experimental run RM 1 in Tables 3-3 and 3-4 for which y1/b

= 7.4, b/d50 = 18.8, V1/Vc = 0.6, and Fr = 0.20. The corresponding values of the dimensionless

parameters in the field for the flood event of July 1994 are y1/b = 6.2, b/d50 = 4800, V1/Vc = 5.94,

and Fr = 0.25. The value of y1/b in the model is slightly higher than in the field in order to obtain

model flow depths large enough for ADV velocity measurements; however, accepted scour

formulas show a negligible effect of this difference in y1/b on scour depths. The corresponding

values of ds/b are 1.04 in the model and 1.07 in the field.

Cross sections and velocity distributions measured in the model and the field are

compared in Fig. 6.14. The initial bed in the laboratory was leveled according to the cross

section measured in March 2001 which is similar to the cross section measured in March 2002

and also in 1980 at the location of the third pier from the left (Station 500 ft) as shown

previously in Fig. 2.15. Most of the pier scour measurements were made at this pier because it

showed the greatest scour due to Tropical Storm Alberto. The bed appears to be at an

Page 134: RP 2002 Final Report - Georgia

121

equilibrium elevation near the top of the upper footing both before Alberto and then several

years after Alberto.

The laboratory model run designated as Run RM 1 in Tables 3-3 and 3-4 was repeated

and both cross sections measured after scour are shown in Fig. 6.14 as Experiments I and II.

They are virtually identical within the experimental uncertainty of scour depth measurement in

the model. When compared to the field measurement made on July 12, 1994, there is remarkable

similarity in the cross sections not only at the pier at Station 500 but also with respect to the

deposition that occurs to the right of this pier around Station 560.

-6

-4

-2

0

2

4

6

8

10

0 100 200 300 400 500 600 700 800

Station, ft

Vel

ocity

, ft/s

40

60

80

100

120

140

160

180

200

Ele

vatio

n, ft

Left (SE) edge channel Right (NW) edge channel

Jul 12, 1994 (Alberto)

Field Data (USGS)Mar 21, 2001

Velocity, ft/s

Velocity, ft/s

Experiment IElevation, ft

2D ADV 2D ADV

EL. 95.24Alberto7/12/94

Experiment IIElevation, ftVelocity, ft/s

Figure 6.14. Comparison of scour cross sections at Flint River bridge from laboratory model and Tropical Storm Alberto.

Page 135: RP 2002 Final Report - Georgia

122

Measured velocities in the model scaled up by the Froude number law are also compared

with field measurements in Fig. 6.14. They are in good agreement not only between the repeated

experiments I and II but also between the laboratory and field measurements. Some discrepancy

can be seen on the right side of the cross section where field velocities are smaller than

laboratory ones, but this may be due to the buildings in the right-side floodplain that were not

included in the model because they did not impact the main pier at Station 500.

The dimensionless maximum pier scour depths are shown in Fig. 6.15 for three

laboratory clear-water scour experiments with differing values of V1/Vc all less than 1.0 but with

constant values of y1/b = 7.4 in rough agreement with the flood event. The values of the Froude

number are shown next to each data point. The data point shown in Fig. 6.15 with a laboratory

Froude number of 0.20 is the one that represents the laboratory results that have been compared

favorably with field data in Fig. 6.14, and it agrees closely with the field scour depth which has a

Froude number of 0.25. Also shown in Fig. 6.15 are the pier scour formulas of Melville (1997)

and Sheppard (2003) for clear-water scour in the laboratory with b/d50 = 18.8 and y1/b = 7.4.

There is relatively good agreement between the laboratory data and these two clear-water scour

formulas. However, as in Fig. 6.13 for the Chattahoochee River, the prototype is in the live-bed

scour regime with d50 = 0.4 mm and b/d50 = 4800, and so the proposed live-bed scour formula of

Sheppard (2003) is also shown in Fig. 6.15 even though it is extrapolated beyond its range. The

results do not agree as well as in Fig. 6.13 perhaps because no attempt has been made to correct

for the effect of the footings.

Page 136: RP 2002 Final Report - Georgia

123

0.0

1.0

2.0

3.0

4.0

5.0

0.5 1.5 2.5 3.5 4.5 5.5 6.5V 1 /V c

d s/ b

River model, b/d50=18.8Sheppard (2003), b/d50=18.8Melville (1997) , b/d50=18.8Field Measurement, b/d50=4813Sheppard (2003), b/d50=4813

0.20Fr=0.25

0.27

0.24

Figure 6.15. Comparison of field and laboratory measurement of scour depths and Sheppard’s equations Flint River at Bainbridge, GA (y1/b = 7.4).

SUMMARY

In this chapter the numerical model has been verified both in terms of detailed velocity

distributions and turbulence kinetic energy. The numerical model has shown that even the

slightest skewness or nonuniformity in the approach flow tends to have a magnified effect on the

turbulence characteristics. Furthermore, the numerical model shows that shear stress alone is not

sufficient to explain the scour patterns and hence the sediment transport rates. The vertical

velocity component before scour, for example, explains much of the measured scour pattern in

the vicinity of a complex pier bent that cannot be accounted for by shear stress alone.

The field and laboratory results in this chapter suggest that in some cases modeling of

live-bed scour might be done in the laboratory in the clear-water regime by preserving Froude

Page 137: RP 2002 Final Report - Georgia

124

number similarity with equality of y1/b values and with b/d50 close to 25. The apparent reduction

in scour at large values of b/d50 is modeled by V1/Vc < 1.0 in the laboratory. However, additional

continuous field measurements such as those obtaned in this project are needed to extend the

predictive range of live-bed scour formulas obtained from laboratory flume data to much larger

values of b/d50 that occur in the field.

Page 138: RP 2002 Final Report - Georgia

125

CHAPTER 7. SUMMARY AND CONCLUSIONS

Field measurements, laboratory modeling, and 3D numerical modeling of scour around bridge

foundations have been successfully combined and applied to specific bridges in Georgia in order

to elucidate the physics of the scouring process and improve scour predictions. In addition,

sediments sampled at the foundations of 10 bridges in Georgia have been characterized with

respect to their erodibility which has been related to easily measured sediment properties. Taken

together, the observations and results in this report will be utilized in Phase 2 of the research to

develop a scour prediction methodology that takes into account site-specific sediment erodibility

properties, time variation of scour depths, scale-up problems of laboratory scour prediction

formulas, and insights about the physics of the flow field around bridge foundations obtained

from the 3D numerical model that can be utilized in one-dimensional modeling.

While the field data collected in Phase 1 of this research have been somewhat limited by

drought conditions, several smaller storm events have been successfully measured at the

Chattahoochee River site and the Ocmulgee River site. The Chattahoochee data obtained for a

bank-full event as well as historical data collected for Tropical Storm Alberto at the Flint River

site have proven to be quite useful in comparisons with the laboratory model data. In addition,

the field data have revealed several important aspects of the bridge scour process including:

• dynamics of live-bed scour around bridge piers in which smaller events gradually fill

remnant scour holes only to be scoured out again by larger storm events;

• simultaneous occurrence of contraction scour and local pier scour and the differentiation

of the two processes based on historical cross sections;

Page 139: RP 2002 Final Report - Georgia

126

• minor cyclical scour and fill during the tidal cycle at the Darien River site with more

significant scour and fill on a seasonal time scale.

The observed simultaneous contraction and local pier scour at the Ocmulgee River site will be

studied more intensely in Phase 2 of this research in the laboratory model to compare field

measurements (mobile and fixed instrumentation), laboratory measurements, and contraction

scour predictions to distinguish between the two types of scour.

The 3D numerical model has revealed a complex flow field around natural river bridge

foundations that is unstable and highly three-dimensional. Large-scale coherent vortex shedding

exists in the vicinity of the bridge foundations with multiple vortices having axes parallel and

perpendicular to the bed. Both the river bathymetry and the presence of multiple pier bents can

influence the flow patterns considerably and need to be taken into account if realistic flow

predictions are to be obtained. The numerical model developed in this research is a powerful

engineering simulation tool for elucidating the complex flow physics of real-life bridge

foundation flows.

The numerical model was successfully validated using velocity and turbulence data from

the laboratory physical models. Using the numerical model it was possible to show that even the

slightest skewness or nonuniformity in the bridge approach flow tends to have a magnified effect

on the turbulence characteristics. Furthermore, the numerical model shows that shear stress alone

is not sufficient to explain the scour patterns and hence sediment transport rates within the scour

zone. The vertical velocity component before scour, for example, explains much of the measured

scour pattern in the vicinity of a complex pier bent that cannot be accounted for by shear stress

alone. Such secondary variables will be related to more conventional hydraulic variables

obtained from the one-dimensional model HEC-RAS in Phase 2 of the research. The insights

Page 140: RP 2002 Final Report - Georgia

127

into the flow structures provided by the 3D numerical model facilitate the understanding and

interpretation of the results from laboratory experiments for scour around bridge foundations and

can provide a link to one-dimensional hydraulic variables that are more easily estimated.

The laboratory model results for scour depth were compared with accepted scour

formulas, and it was found that the Sheppard and the Melville clear-water scour formulas

provide reasonable agreement with the laboratory data except at large values of the Froude

number at which the HEC-18 formula performs better. These comparisons bring into focus the

laboratory scale-up problem in which it is apparent that proper selection of the sediment size in

the model is extremely important in order to avoid violation of similarity requirements due to the

Froude number (Fr) becoming too large and the value of the ratio of pier width to sediment

diameter (b/d50) becoming too small in the laboratory.

A laboratory modeling strategy was applied to the river models in which the sediment

size was chosen to reproduce the field value of the Froude number as closely as possible in the

clear-water scour regime while maintaining the value of b/d50 as close to 25 as possible at which

it has little influence on scour depth. This strategy was proven to be successful both for a bank-

full event measured in this study at the Chattahoochee River site and an extreme historical flood

event at the Flint River site. Very good comparisons were obtained not only for maximum pier

scour depth but also for cross-sectional changes and measured velocity distributions at the

bridge. The field and laboratory results suggest that reproduction of live-bed scour scour depths

may be achieved in the laboratory in the clear-water regime by preserving Froude number

similarity with equality of y1/b values and with b/d50 close to 25. The apparent reduction in scour

at large values of b/d50 is modeled by V1/Vc < 1.0 in the laboratory. However, additional

continuous field measurements such as those obtained in this project are needed to extend the

Page 141: RP 2002 Final Report - Georgia

128

predictive range of live-bed scour formulas obtained from laboratory flume data to much larger

values of b/d50 that occur in the field. An expected outcome of this effort is that predictions of

scour for smaller bridges should follow the same formulas as larger bridges if the scaling

relationships are properly developed.

The measurement of sediment erodibility in the scour flume was successful, but the

degreee of stratification of sediment properties was greater than expected. While ranges of

erodibility parameters such as critical shear stress and erosion rate constants can be defined for

physiographic regions in Georgia, they are by no means universal but nevertheless can be useful

in providing guidance in the initial bridge foundation design to prevent possible scour failure.

For a more definitive measure of erosional strength, the critical shear stress was successfully

correlated with percent fines and median sediment grain diameter as an approximate measure of

the influence of interparticle forces on the erodibility of sediments consisting of both fine and

coarse-grained fractions. This is considered to be an important step forward in characterizing

sediment resistance to scour in scour-prediction procedures. For high-risk bridges, site-specific

tests for scour resistance such as those conducted in this study are recommended.

Based on the combined field, laboratory, and 3D numerical results obtained to date, it is

clear that a scour-prediction methodology needs to incorporate the effects of sediment scour

resistance for mixed grain sizes, time variation of scour depth and resultant sediment transport

rates out of the scour hole, explanatory hydraulic variables driving the scour process as identified

by the numerical model both before and after scour, live-bed scour processes observed in the

field, and an understanding of laboratory scale-up issues. Much of the information needed to

develop such a methodology has been obtained in the Phase 1 research, but the Phase 2 research

is expected to add needed insights into contraction scour at the Ocmulgee River site, tidal

Page 142: RP 2002 Final Report - Georgia

129

processes at the Darien River and Altamaha River sites, and values of one-dimensional hydraulic

parameters after scour as well as before scour to be incorporated into scour prediction formulas.

The Phase 2 research will also provide invaluable and essential field data to validate and extend

the concepts developed thus far in this study.

Page 143: RP 2002 Final Report - Georgia

130

REFERENCES

Alhadeff, S. J., Musser, J. W., Sandercock, A. C., and Dyar, T. R. (2000). “Digital Environmental Atlas of Georgia.” Georgia Geologic Survey Publication CD-1. U. S. Geological Survey and Georgia Department of Natural Resources. 2 Disks. Briaud, J. L., Ting, C. K., Chen, H.C., Gudavalli R., Perugu, S., and Wei, G. (1999). “SRICOS: Prediction of scour rate in cohesive soils at bridge piers,” J. Geotechnical and Environmental Eng., ASCE, 125(4), 237-246. Briaud, J. L., Ting, C. K., Chen, H.C., Cao, Y., Han, S. W., and Kwak, K. W. (2001). “Erosion function apparatus for scour rate prediction.” J. Geotechnical and Environmental Eng., ASCE, 127(2), 105-113. Cardoso, A. H., and Bettess, R. (1999). “Effects of time and channel geometry on scour at bridge abutments.” J. Hydr. Engrg., ASCE, 125(4), 388-399. Dargahi, B. (1990), “Controlling mechanism of local scour,” J. Hydr. Engrg., ASCE, 116(10), 1197-1214. Dargahi, B. (1989), “The turbulent flow field around a circular cylinder,” Exps. Fluids 8, 1-12. Dennett, K. E., Sturm, T.W., Amirtharajah, A., and Mahmood, T. (1998), "Effect of adsorbed natural organic matter on the erosion of kaolinite sediments," Water Environment Research, 70(3), 268-275. Dou, X., Jia, Y., and Wang, S. Y. (1996), “Numerical simulation of bridge abutment scour development,” Proc. Of the 1996 North American Water and Environment Congress, ASCE, Reston, VA. Ettema, R. (1980). “Scour at bridge piers.” Rep. No. 216, School of Engrg., University of Auckland, Auckland, New Zealand. Ettema, R., Melville, B. W., and Barkdoll, B. (1998). “Scale effect in pier-scour experiments.” J. Hyd. Engrg, ASCE, 124(6), 639-642. Hoepner, M. A. (2001). “Stability of cohesive sediments from flume and rheometer measurements.” M. S. Thesis, Georgia Institute of Technology, Atlanta, GA. Julien, P. Y. (1995), “Erosion and Sedimentation.” Cambridge University Press, NY, 280 pp. Kamphuis, J. W., and Hall, K. R. (1983). “Cohesive material erosion by unidirectional current.” J. Hydr. Engrg., ASCE, 109(1), 49-61. Kandiah, A. (1974). “Fundamental aspects of surface erosion of cohesive soils.” Ph. D. Thesis, University of California, Davis, Calif.

Page 144: RP 2002 Final Report - Georgia

131

Krone, R. B. (1999). “Effects of bed structure on erosion of cohesive sediments.” J. Hydr. Engrg., ASCE, 125(12), 1297-1301. Landers, M. N. and Mueller, D. S. (1996). “Evaluation of selected pier-scour equations using field data.” Transp. Res. Record, National Research Council, Washington, D.C., 186-195. Lane, S. N., Biron, P. M., Bradbrook, K. F., Butler, J. B., Chander, J. H., Crowell, M. D., McLelland, S. J., Richards, K. S., and Roy, A. G. (1998). “Three-dimensional measurement of river channel flow processes using acoustic Doppler velocimetry.” Earth Surface Processes and Landforms, 23, 1247-1267. Lee, S. C., Mehta, A. J., and Parchure, T.M., (1994). “Cohesive sediment erosion: Part I, test devices and field instrument assemblies. Part II, relationship between the erosion rate constant and bed shear strength.” Internal Report, Coastal and Oceanographic Engineering Dept., University of Florida, Gainesville, Fla. Liang, G., Lee, SeungOh, Sotiropoulos, F., and Sturm, T. “3D usteady RANS modeling of complex hydraulic flows. II. Model validation” Submitted to J. Hydr. Engrg., ASCE, 2004. Lin, F. B. and Sotiropoulos, F. (1997), “Assessment of artificial dissipation models for three-dimensional incompressible flow solutions,” ASME Journal of Fluids Engineering, 119(2), 331-340. Mahmood, T., Amirtharajah, A., Sturm, T. W., and Dennet, K. E. (2001). “A micromechanics approach for attachment and detachment of asymmetric particles.” Colloids and Surfaces: A. Physicochemical and Engineering Aspects, 177, 99-110. Martin, V., Fisher, T. S. R., Millar, R. G., and Quick, M. C. (2002). “ADV data analysis for turbulent flows: low correlation problem.” Hydraulic Measurements and Experimental Methods 2002, Proc. Of the Specialty Conf., ed. By T. L. Wahl, C. A. Pugh, K. A. Oberg, and T. B. Vermeyen, ASCE, Reston, VA. McNeil, J., Taylor, C., and Lick, W. (1996). “Measurements of erosion of undisturbed bottom sediments with depth.” J. Hydr. Engrg., ASCE, 122(6), 316-324. Melville, B. W. (1997). “Pier and abutment scour: integrated approach.” J. Hydr. Engrg, ASCE, 123(2), 125-136. Melville, B. W., and Chiew, Y. M. (1999). “Time scale for local scour at bridge piers.” J. Hydr. Engrg., ASCE, 125(1), 59-65. Melville, B. W., and Coleman, S. E. (2000). Bridge Scour. Water Resources Publications, Highlands Ranch, Colorado

Page 145: RP 2002 Final Report - Georgia

132

Mehta, A. J. (1991). “Characterization of cohesive soil bed surface, with special reference o the relationship between erosion shear strength and bed density,” Report No. UFL/COEL.MP-91/4, Coastal and Oceanographic Engineering Dept., University of Florida, Gainesville, Fla. Morris, J. L., and Pagan-Ortiz, J. E. (1999). “Bridge scour evaluation program in the United States.” Stream Stability and Scour at Highway Bridges, E.V. Richardson and P.F. Lagasse,eds., ASCE,61-70. Navarro, Hernan Ricardo (2004). “Flume measurements of erosion characteristics of soils at bridge foundations in Georgia,” M. S. Thesis, Georgia Institute of Technology, Atlanta, GA. Neary, V. S., Sotiropoulos, F., and Odgaard, A. J. (1999) “Three-dimensional numerical model of lateral-intake flows,” J. Hydr. Engrg., ASCE, 125(2), pp. 126-140. Nezu, I., and Nakagawa, H. (1993). Turbulence in open channel flows. A.A. Balkema, Rotterdam, The Netherlands. Nezu, I., and Rodi, W. (1986). “Open channel flow measurements with a laser Doppler anemometer,” J. Hydr. Engrg., ASCE, 112(5), 335-355. Nikora, V., and Goring, G. (2000). “Flow turbulence over fixed and weakly mobile gravel beds.” J. Hydr. Engrg., ASCE, 126, 9, 679-690. Olsen, N. R. B., and Melaaen, M. C. (1993), “Three-dimensional calculation of scour around cylinders,” J. Hydr. Engrg., ASCE, 119(9), pp. 1048-1054. Parola, A. C., Hagerty, D. J., Mueller, D. S., Melville, B. W., Parker, G., and Usher, J. S. (1997b), “The need for research on scour at bridge crossings,” Proc. Of XXVII IAHR Congress, Managing Water: Coping with Scarcity and Abundance, San Francisco, CA, August 10-15, 1997, pp. 124-129. Partheniades, E. (1965). “Erosion and deposition of cohesive soils,” J. Hydr. Engrg., ASCE, 91(1), 105-139. Raudkivi, A. J. (1986), “Functional trends of scour at bridge piers,” J. Hydr. Engrg., ASCE, 112(1), 1-13. Ravens, T. M., and Gschwend, P. M., (1999). “Flume measurements of sediment erodibility in Boston Harbor,” J. Hydr. Engrg., ASCE, 125(10), 998-1005. Ravisangar, V., Dennet, K. E., Sturm, T. W., and Amirtharajah, A. (2001). “Effect of sediment pH on resuspension of kaolinite sediments,” J. Environ. Engrg.,ASCE, 127(6), 531-538. Richardson, E. V., and Davis, S. R. (2001). “Evaluating scour at bridges” Report HEC-18, 4th Edition, U.S. Federal Highway Administration, May.

Page 146: RP 2002 Final Report - Georgia

133

Santamarina, J. C., Klein, K. A., Fam, M. A. (2001). Soils and waves. John Wiley and Sons Ltd., England, 488 pp. Sheppard, D. Max (2003). “Large scale and live bed local pier scour experiments.” Final Report, Florida Department of Transportation, Tallahassee, BB-473. Sinha, S. K., Sotiropoulos, F., and Odgaard, A. J. (1998), “Three-dimensional numerical model of turbulent flow through natural rivers of complex bathymetry,” J. Hydr.Engrg., ASCE, 124(1), pp. 13-24. SonTek (2001). “Acoustic Doppler velocimeter (ADV) principles of operation.” SonTek Technical Notes, SonTek, San Diego, CA. Sotiropoulos, F., Chrisohoides, A., and Sturm, T. W. (1999), “prediction of bridge-abutment flows using advanced turbulence models,” proc. of the 1999 ASCE Int. Water Resources Engineering Conf., Seattle, WA, August 1999. Sotiropoulos, F., and Patel, V. C. (1994), "Prediction of turbulent flow through a transition duct using a second-moment closure," AIAA J., 32, No. 11, pp. 2194-2204. Sotiropoulos, F., and Patel, V. C. (1995a), "On the role of turbulence anisotropy and near-wall modeling in predicting complex, 3D shear flows," AIAA Journal, 33, No. 3, pp. 504-514. Sotiropoulos, F., and Patel, V. C. (1995b), "Application of reynolds-stress transport models to stern and wake flows," J. of Ship Research, Vol. 39, No. 4, pp. 263-283. Sotiropoulos, F., and Ventikos, Y. (1998) “Flow Through a Curved Duct Using Linear and Non-Linear Two-Equation Turbulence Models,” AIAA Journal 36(7), pp. 1256-1262. Stamey, Timothy C. (1996), "Summary of Data-collection activities and effects of flooding from Tropical Storm Alberto in parts of Georgia, Alabama, and Florida, July 1994," U.S. Geological Survey Open-File Report 96-228, 23p. Steger, J. L. and Benek, J. A. (1987), “On the use of composite grid schemes in computational aerodynamics,” Computer Methods in Applied Mechanics and Engineering, 64(1-3), 301-320. Sturm, T. W., and Janjua, N. S. (1994), “Clear-water scour around abutments in floodplains,” J. Hydr. Engrg., ASCE, 120(8), pp. 956-972. Sturm, T. W. and Sadiq, Aftab (1996), "Clear-water scour around bridge abutments under backwater conditions," Transportation Research Record 1523, TRB, National Research Council, Washington, D.C., pp. 196-202. Sturm, T. W. (1998), "Effect of compound channel hydraulics on abutment scour," Final Report, GDOT Project 9411, Office of Materials and Research, Georgia Dept. Of Transportation.

Page 147: RP 2002 Final Report - Georgia

134

Sturm, T. W. (1999a), "Abutment scour in compound channels," Stream Stability and Scour at Highway Bridges, ASCE Scour Compendium, pp. 443-456. Sturm, T. W. (1999b), "Abutment scour studies for compound channels," Report No. FHWA-RD-99-156, Federal Highway Administration, U. S. Dept. of Transportation. Sturm, T. W. (2001). Open Channel Hydraulics. McGraw-Hill, New York. Tang, H. S., Jones, S. C., and Sotiropoulos, F. (2003), “An overset-grid method for 3D unsteady incompressible flows,” Jounral of Computational Physics, 191(2), 567-600. Voulgaris, G., and Trowbridge, J. H. (1998). “Evaluation of the acoustic Doppler velocimeter (ADV) for turbulence measurements.” J. Atmos. And Oceanic Technology, American Meteorological Society, 15, 272-289. Wahl, T. (2002). “Analyzing ADV data using WinADV.” Hydraulic Measurements and Experimental Methods 2002, Proc. Of the Specialty Conf., ed. By T. L. Wahl, C. A. Pugh, K. A. Oberg, and T. B. Vermey Zreik, D.A., Krishnappan, B. G., Germaine, J. T., Madsen, O. S., and Ladd, C. C. (1998). “Erosional and mechanical strengths of deposited cohesive sediments.” J. Hydr. Engrg.,ASCE, 124(11), 1076-1085.

Page 148: RP 2002 Final Report - Georgia

135

APPENDIX A

EROSION AND SOIL PROPERTY TEST RESULTS

Page 149: RP 2002 Final Report - Georgia

136

Table A-1. Linear, piecewise linear and exponential critical shear stress values for the erosion rate vs. applied shear stress models [(1) = lower line segment, (2) = upper line segment]. Linear Exponential

−⋅=

c

cMEτ

ττ

−⋅

⋅= c

ca

c eEE τττ

Critical Erosion = 0.00190 kg/m2/s τc, Pa s.e. τc R2 τc, Pa s.e. τc R2

Sand d50=1.16mm (1) 1.03 0.15 0.95 0.38 0.20 0.95 Sand d50=1.16mm (2) 2.91 0.04 0.99 Peachtree Creek 2.18 0.43 0.92 Murray 3'-4' (1) 4.05 0.50 0.75 3.14 0.63 0.81 Murray 3'-4' (2) 5.33 0.29 0.44 Towns 5'-5'6" 17.21 0.48 0.91 17.02 0.26 0.98 Towns 6'-7' (1) 11.31 1.44 0.65 12.55 0.95 0.68 Towns 6'-7' (2) 15.62 0.48 0.84 Towns 7'-8' 6.82 0.39 0.72 4.96 0.62 0.89 Habersham 11'-12' 17.35 0.84 0.57 17.57 0.78 0.56 Habersham 12'-12'6" 3.29 0.32 0.79 1.75 0.55 0.85 Habersham 12'6"-14" 4.54 0.28 0.85 4.35 0.38 0.77 Haralson 12'-15'5" (1) 5.77 0.27 0.70 4.38 0.33 0.89 Haralson 12'-15'5" (2) 8.68 0.33 0.81 Wilkinson 36'6"-37'6" 0.44 0.14 0.98 Bibb 25'-25'8" 9.68 0.29 0.79 10.08 0.31 0.66 Bibb 30'8"-31'4" 3.32 0.34 0.73 2.36 0.54 0.79 Bibb 31'4"-32' (1) 5.11 1.00 0.65 6.32 0.56 0.85 Bibb 31'4"-32' (2) 9.45 0.98 0.76 Effingham 21'-22' 3.24 0.25 0.81 Decatur 20'-20'4" (1) 7.90 0.36 0.96 7.98 0.17 0.99 Decatur 20'-20'4" (2) 9.88 0.32 0.95 McIntosh 10'-12' 17.17 0.15 0.94 15.92 0.54 0.88 Average 0.45 0.80 0.57 0.84

Page 150: RP 2002 Final Report - Georgia

137

Table A-2. Linear, piecewise linear and exponential slope coefficients for the erosion rate vs. applied shear stress models [(1) = lower line segment, (2) = upper line segment]. Linear Exponential

−⋅=

c

cMEτ

ττ

−⋅

⋅= c

ca

c eEE τττ

Critical Erosion = 0.00190 kg/m2/s

M kg/m2/s s.e. M s.e./M a s.e. a s.e./a

Sand d50=1.16mm (1) 0.063 0.007 11% 0.626 0.053 8% Sand d50=1.16mm (2) 1.81 0.11 6% Peachtree Creek 0.410 0.083 20% Murray 3'-4' (1) 0.209 0.084 40% 5.12 1.45 28% Murray 3'-4' (2) 2.45 1.95 80% Towns 5'-5'6" 0.096 0.030 32% 12.7 1.9 15% Towns 6'-7' (1) 0.022 0.009 43% 7.6 2.6 34% Towns 6'-7' (2) 0.41 0.18 44% Towns 7'-8' 3.29 1.44 44% 8.2 2.0 25% Habersham 11'-12' 0.043 0.017 39% 8.3 3.3 40% Habersham 12'-12'6" 0.309 0.080 26% 2.37 0.49 21% Habersham 12'6"-14" 0.021 0.002 10% 2.33 0.30 13% Haralson 12'-15'5" (1) 0.117 0.023 20% 3.5 0.33 9% Haralson 12'-15'5" (2) 1.60 0.38 24% Wilkinson 36'6"-37'6" 0.010 0.001 11% Bibb 25'-25'8" 0.053 0.011 21% 11.0 3.2 29% Bibb 30'8"-31'4" 0.084 0.029 35% 2.74 0.82 30% Bibb 31'4"-32' (1) 0.007 0.002 33% 2.08 0.33 16% Bibb 31'4"-32' (2) 0.048 0.019 40% Effingham 21'-22' 2.53 0.87 34% Decatur 20'-20'4" (1) 0.033 0.007 21% 6.2 0.41 7% Decatur 20'-20'4" (2) 0.192 0.043 23% McIntosh 10'-12' 2.14 0.38 18% 28.7 7.6 26% Average 29% 22%

Page 151: RP 2002 Final Report - Georgia

138

Table A-3. Results of erosion tests and soil property tests on Murray County sample.

Site Murray 1'-2' Murray 2'-3' Murray 3'-4' Critical Shear Stress (Pa) >21 <3 4.05

Sediment SC-SM SP-SM SM

Group Name Silty, Clayey Sand Poorly Graded Sand with Silt Silty Sand

Color Light Brown Gray Light Gray Dry Density (Kg/m3) 1693 1695 1649

e (void ratio) 0.56 0.55 0.59 Bulk Density (Kg/m3) 1993 1963 2220

Water Content 18% 16% 35% Specific Gravity 2.64 2.62 2.63 Organic Matter 3.1% 1.6% 2.4% Liquid Limit 22% NP NP Plasticity Index 5% NP NP d50 (mm) 0.0802 0.6734 0.3112 Sand 55% 90% 75% Silt 31% 7% 18% Clay 15% 3% 7%

Page 152: RP 2002 Final Report - Georgia

139

Table A-4. Results of erosion tests and soil property tests on Towns County sample.

Site Towns 5' - 5' 6"

Towns 5' 6" - 6' Towns 6' – 7' Towns 7' - 8'

Critical Shear Stress (Pa) 17.21 >21 11.31 6.82

Sediment ML MH ML SP

Group Name Sandy Silt Elastic Silt with Sand Sandy Silt

Poorly Graded Sand with

Gravel Color Gray Brown Gray Brown Gray Brown Light Brown Dry Density (Kg/m3) 1099 876 1019 1588

e (void ratio) 1.44 2.03 1.68 0.71 Bulk Density (Kg/m3) 1477 1177 1369 2079

Water Content 34% - 34% 31%

Specific Gravity 2.68 2.65 2.73 2.71

Organic Matter 3.6% 2% 2% 1%

Liquid Limit 44% 51% 41% NP Plasticity Index 12% 13% 7% NP

d50 (mm) 0.032 0.020 0.047 1.19 Sand 44% 37% 50% 97% Silt 36% 39% 33% 3% Clay 20% 24% 17% 0%

Page 153: RP 2002 Final Report - Georgia

140

Table A-5. Results of erosion and soil property tests on Habersham County sample.

Site Habersham 10' - 11'

Habersham 11' - 12'

Habersham 12' – 12' 6"

Habersham 12' 6" -

14'

Habersham 20' - 21'

6" Critical Shear Stress (Pa) >21 17.35 3.29 4.54 ~2.5

Sediment ML ML SM SM SP-SM

Group Name Sandy Silt Sandy Silt Silty Sand Silty Sand

Poorly Graded

Sand with Silt

Color Light Brown Light Brown Tan Tan Gray

Dry Density (Kg/m3) 1410 1473 1366 1463 1586

e (void ratio) 0.89 0.91 0.98 0.81 0.71 Bulk Density (Kg/m3) 1819 1909 1678 1893 1962

Water Content 29% 30% 23% 29% 24% Specific Gravity 2.66 2.81 2.71 2.65 2.71

Organic Matter 5.5% 4% 2% 2% 3% Liquid Limit 35% 37% NP NP NP Plasticity Index 11% 11% NP NP NP d50 (mm) 0.031 0.043 0.153 0.163 0.265 Sand 46% 48% 78% 83% 92% Silt 32% 32% 14% 11% 7% Clay 22% 20% 8% 5% 1%

Page 154: RP 2002 Final Report - Georgia

141

Table A-6. Results of erosion tests and soil property tests on Haralson County sample.

Site Haralson 12’- 15’5”

Haralson 15’5”-15’7”

Haralson 15’7"-17’

Critical Shear Stress (Pa) 5.77 >12 ~3

Sediment SM SM ML Group Name Silty Sand Silty Sand Sandy Silt Color Mustard Yellow Yellow Orange Tan Dry Density (Kg/m3) 1638 1843 -

e (void ratio) 0.63 0.45 - Bulk Density (Kg/m3) 2026 2279 -

Water Content 24% - - Specific Gravity 2.67 2.68 2.82 Organic Matter 2.0% 1.3% 3.0% Liquid Limit 35% 29% 32% Plasticity Index 9% 5% 4% d50 (mm) 0.27 0.44 0.04 Sand 71% 69% 48% Silt 27% 28% 51% Clay 3% 3% 1%

Page 155: RP 2002 Final Report - Georgia

142

Table A-7. Results of erosion tests and soil property tests on Bibb County sample.

Site Bibb 25' - 25' 8"

Bibb 25' 8" – 27'

Bibb 30' - 30' 8"

Bibb 30' 8" - 31' 4"

Bibb 31' 4" - 32'

Critical Shear Stress (Pa) 9.68 ~2.5 ~16.5 3.32 5.11

Sediment CL SM SC SP-SM ML

Group Name Lean Clay with Sand Silty Sand Clayey

Sand

Poorly Graded

Sand with Silt

Sandy Silt

Color Light Brown Gray Light

Brown Light

Brown Dark Gray

Dry Density (Kg/m3) 1261 1631 1596 1316 1162

e (void ratio) 1.07 0.62 0.68 1.02 1.19 Bulk Density (Kg/m3) 1749 1949 1973 1715 1513

Water Content 39% 20% 24% 30% Specific Gravity 2.62 2.64 2.69 2.66 2.54

Organic Matter 10.3% 6.5% 6.7% 6.2% 16.4% Liquid Limit 36% NP 32% NP 39% Plasticity Index 18% NP 11% NP 12% d50 (mm) 0.0074 0.250 0.111 0.159 0.036 Sand 26% 84% 59% 90% 44% Silt 40% 9% 20% 4% 33% Clay 34% 7% 21% 6% 23%

Page 156: RP 2002 Final Report - Georgia

143

Table A-8. Results of erosion tests and soil property tests on Wilkinson County sample.

Site Wilkinson 36'6”-37'6” Wilkinson 37'6”-38'6” Critical Shear Stress (Pa) 0.44 >21 Sediment SP-SM CH

Group Name Poorly Graded Sand with Silt Fat Clay

Color Tan Light Brown Dry Density (Kg/m3) 1544 1657 e (void ratio) 0.70 0.58 Bulk Density (Kg/m3) 2007 2227 Water Content 30% 34% Specific Gravity 2.63 2.61 Organic Matter 0.3% 6.6% Liquid Limit NP 51% Plasticity Index NP 23% d50 (mm) 0.1803 0.004 Sand 93% 11% Silt 3% 49% Clay 4% 40%

Page 157: RP 2002 Final Report - Georgia

144

Table A-9. Results of erosion tests and soil property tests on Effingham County sample.

Site Effingham 20’-21’ Effingham 21’-22’ Critical Shear Stress (Pa) >21 3.24 Sediment SC SP Group Name Clayey Sand Poorly Graded Sand Color Gray Light Gray Dry Density (Kg/m3) 1430 - e (void ratio) 0.78 - Bulk Density (Kg/m3) 1733 - Water Content 21% 21% Specific Gravity 2.54 2.64 Organic Matter 2.2% 0.0% Liquid Limit 36% NP Plasticity Index 19% NP d50 (mm) 0.3 0.45 Sand 67% 99% Silt 15% 1% Clay 18% 0%

Page 158: RP 2002 Final Report - Georgia

145

Table A-10. Results of erosion tests and soil property tests on Decatur County sample.

Site Decatur 20'-20'4" Decatur 20'4"-22' Critical Shear Stress (Pa) 7.90 ~2.5 Sediment SC SM Group Name Clayey Sand Silty Sand Color Red Brown Brown Dry Density (Kg/m3) 1761 1548 e (void ratio) 0.49 0.71 Bulk Density (Kg/m3) 2114 1887 Water Content 20% 22% Specific Gravity 2.62 2.65 Organic Matter 4.8% 0.6% Liquid Limit 28% NP Plasticity Index 12% NP d50 (mm) 0.131 0.404 Sand 60% 83% Silt 9% 3% Clay 31% 14%

Page 159: RP 2002 Final Report - Georgia

146

Table A-11. Results of erosion tests and soil property tests on Berrien County sample.

Site Berrien 25'-25'6"

Berrien 25'6"-27'

Berrien 30'-30'6"

Berrien 30'6"-32'

Critical Shear Stress (Pa) >21 >21 >21 >21

Sediment CH SC SM CH

Group Name Sandy Fat Clay Clayey Sand Silty Sand Fat Clay with Sand

Color Gray Tan Light Gray Brown Gray Dry Density (Kg/m3) 1065 1246 1482 1052

e (void ratio) 1.55 1.05 0.74 1.56 Bulk Density (Kg/m3) 1698 1895 1764 1716

Water Content 59% 52% 19% 63%

Specific Gravity 2.72 2.55 2.58 2.7

Organic Matter 4.6% 6.1% 5.6% 4.3%

Liquid Limit 103% 76% 22% 114% Plasticity Index 72% 39% 2% 69%

d50 (mm) <0.001 0.084 0.144 <0.001 Sand 38% 53% 72% 26% Silt 5% 23% 10% 6% Clay 57% 24% 18% 68%

Page 160: RP 2002 Final Report - Georgia

147

Table A-12. Results of erosion tests and soil property tests on McIntosh County sample.

Site McIntosh 10’-12’ Critical Shear Stress (Pa) 17.17 Sediment SC Group Name Clayey Sand with Gravel (Shells) Color Black Dry Density (Kg/m3) 1298 e (void ratio) 1.00 Bulk Density (Kg/m3) 1728 Water Content 33% Specific Gravity 2.6 Organic Matter 5.7% Liquid Limit 32% Plasticity Index 16% d50 (mm) 1 Sand 87% Silt 6% Clay 7%