interaction of bottom turbulence and cohesive sediment … · interaction of bottom turbulence and...

103
INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT ON THE MUDDY ATCHAFALAYA SHELF, LOUISIANA, USA By ILGAR SAFAK A DISSERTATION PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY UNIVERSITY OF FLORIDA 2010

Upload: letruc

Post on 17-Aug-2018

243 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT ON THEMUDDY ATCHAFALAYA SHELF, LOUISIANA, USA

By

ILGAR SAFAK

A DISSERTATION PRESENTED TO THE GRADUATE SCHOOLOF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT

OF THE REQUIREMENTS FOR THE DEGREE OFDOCTOR OF PHILOSOPHY

UNIVERSITY OF FLORIDA

2010

Page 2: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

c© 2010 Ilgar Safak

2

Page 3: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

To Besiktas

3

Page 4: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

ACKNOWLEDGMENTS

I would like to thank my advisor Dr. Alex Sheremet to give me the opportunity to

work with him at the University of Florida. His expertise, readiness for advising, and

suggestions made everything much easier for me. I have also learned from him the real

scientific approach to all sorts of problems, including those in daily life. For me, this is as

valuable as the research work.

This research was supported by Office of Naval Research funding of Contracts No.

N00014-07-1-0448 and N00014-07-1-0756.

Dr. Tian-Jian Hsu from the University of Delaware is greatly acknowledged for

agreeing to share with me the boundary layer model he developed. My research has

benefited substantially from this model and through his guidance. He was initially an

internal member in my committee, but unfortunately he had to be taken out, as he left

the University of Florida.

Special thanks go to Sergio Jaramillo, who is now in the University of Hawaii at

Manoa, and especially Bilge Tutak, who have never hesitated to leave their works aside

to be able to give me a hand for my work.

Dr. Mead A. Allison from the University of Texas, Austin provided extremely valuable

data, help, and guidance at every level of this study.

My committee members Dr. Arnoldo Valle-Levinson, Dr. Jane Mckee Smith,

Dr. Peter N. Adams, and Dr. Donald N. Slinn spent their valuable time to evaluate

the progress of my research work. Dr. Valle-Levinson is further acknowledged for

introducing me some oceanographic concepts that I was not very familiar with.

The data set, used in this study, was collected during a field experiment that

benefited from the efforts of Viktor B. Adams, Sidney Schofield and Jimmy Joiner from

our Coastal Engineering Laboratory, Daniel D. Duncan from the University of Texas,

Austin, and the field support group of LUMCON. Sergio Jaramillo, Uriah M. Gravois, and

Jungwoo Lee also assisted in the deployment and retrieval of the instruments.

4

Page 5: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

Dr. Clinton D. Winant from Scripps Institution of Oceanography spent his valuable

time on tutoring me and the other graduate students towards our Ph.D. qualifying tests in

2007.

While it is not in the scope of this dissertation, I have had the opportunity to work

on dissipation of surface wave energy while propagating over muddy seafloors, using

a comprehensive data set that was kindly provided by Dr. Steve Elgar and Dr. Britt

Raubenheimer from Woods Hole Oceanographic Institution.

Tracy J. Martz, Chelsea L. Sydow, and Chloe D. Winant kindly agreed to proofread

this dissertation.

I would like to thank my dear friend Hande Caliskan, who graduated from our

program in 2006, and Dr. Aysen Ergin, my advisor during my M.S. studies at Middle

East Technical University. Dr. Ergin made me contact with Hande for applying to

University of Florida, and Hande’s guidance throughout the entire application process

helped substantially.

Although being far away from me, knowing that Irmak Yesilada, Ozan Gokler and

Murat Dilman - the co-starrings if a desperate director would make a biographical sketch

of my life one day- are somewhere in this world always helps me to enjoy life, and get

along with all sorts of difficulties easily.

Cihat and Sukran, you GUYS are the best!

5

Page 6: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

TABLE OF CONTENTS

page

ACKNOWLEDGMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

ABSTRACT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

CHAPTER

1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

1.1 Cohesive Sediments in Marine Environment . . . . . . . . . . . . . . . . . 131.2 Turbulence in Combined Wave-Current Flow . . . . . . . . . . . . . . . . 161.3 Interaction of Turbulent Flow with Cohesive Sediments . . . . . . . . . . . 171.4 Objectives of This Study . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

2 FIELD EXPERIMENT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

2.1 Experiment Site . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242.2 Instrumentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 272.3 General Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

3 DATA ANALYSIS METHODS . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

3.1 Wave Spectral Calculations . . . . . . . . . . . . . . . . . . . . . . . . . . 393.2 Estimation of Reynolds Stresses in the Presence of Surface Waves . . . 42

3.2.1 Definition of Reynolds stress . . . . . . . . . . . . . . . . . . . . . 423.2.2 Wave bias in Reynolds stress estimates . . . . . . . . . . . . . . . 433.2.3 Two-sensor methods to reduce wave bias . . . . . . . . . . . . . . 45

3.3 Logarithmic Law of the Wall . . . . . . . . . . . . . . . . . . . . . . . . . . 49

4 OBSERVATIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

4.1 General Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 534.2 The LISST Data Set from 2006 . . . . . . . . . . . . . . . . . . . . . . . . 56

5 BOTTOM BOUNDARY LAYER MODELING . . . . . . . . . . . . . . . . . . . . 67

5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 675.2 Governing Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

5.2.1 Momentum balance . . . . . . . . . . . . . . . . . . . . . . . . . . 685.2.2 Sediment concentration balance . . . . . . . . . . . . . . . . . . . 705.2.3 Turbulent kinetic energy balance and balance of turbulent kinetic

energy dissipation rate . . . . . . . . . . . . . . . . . . . . . . . . . 705.2.4 Sediment definition . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

6

Page 7: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

5.3 Analytical Flocculation Model . . . . . . . . . . . . . . . . . . . . . . . . . 725.4 Application . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 745.5 Model Sensitivity to Floc Size . . . . . . . . . . . . . . . . . . . . . . . . . 77

6 CONCLUSION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84

APPENDIX

A DIRECT ESTIMATION OF REYNOLDS STRESSES . . . . . . . . . . . . . . . 88

REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95

BIOGRAPHICAL SKETCH . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103

7

Page 8: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

LIST OF TABLES

Table page

2-1 Location, mean depth, and retrieval dates of the instrumented platforms . . . . 32

5-1 Numerical coefficients in the k − ε closure . . . . . . . . . . . . . . . . . . . . . 79

8

Page 9: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

LIST OF FIGURES

Figure page

1-1 Noncohesive sand grains and cohesive fluid mud. . . . . . . . . . . . . . . . . 22

1-2 Comparison of spectral evolution of waves over muddy and sandy seafloors. . 23

2-1 Gulf of Mexico . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

2-2 Bathymetry of the Atchafalaya Shelf and the locations of the instrumentedplatforms. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

2-3 A typical instrumentation platform . . . . . . . . . . . . . . . . . . . . . . . . . . 35

2-4 The synchronized ADVs and the OBS-5 on the platform before the deployment 36

2-5 Configuration of the instrument array . . . . . . . . . . . . . . . . . . . . . . . . 37

2-6 Wind, wave, and near-bed conditions on the Atchafalaya Shelf throughout the2008 experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

3-1 Examples of ogive curves as a function of the dimensionless wavenumber . . . 52

4-1 Tidal variations during the experiment . . . . . . . . . . . . . . . . . . . . . . . 58

4-2 A one-minute segment of the flow velocity, recorded by the ADV array . . . . . 59

4-3 Wind, wave, and current conditions on the Atchafalaya Shelf throughout the2-week experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

4-4 Bulk spectral characteristics of waves, and the measurements of suspendedsediment concentration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

4-5 Reynolds stress estimates versus the quadratic drag relation . . . . . . . . . . 62

4-6 Reynolds stress estimates, and the measurements of suspended sedimentconcentration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

4-7 Wave and current observations during the 2006 experiment . . . . . . . . . . . 64

4-8 Examples of logarithmic layer fits, and grain size distributions from the 2006experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

4-9 Observations of particle size distribution and wave-turbulence conditions in2006 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

5-1 An example of time-series in the model simulations . . . . . . . . . . . . . . . . 80

5-2 Comparison of the observations and the model results . . . . . . . . . . . . . . 81

5-3 Analysis of the model representation of the turbulent kinetic energy balance . . 82

9

Page 10: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

5-4 Effect of varying floc size D on the model calculations . . . . . . . . . . . . . . 83

10

Page 11: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

Abstract of Dissertation Presented to the Graduate Schoolof the University of Florida in Partial Fulfillment of theRequirements for the Degree of Doctor of Philosophy

INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT ON THEMUDDY ATCHAFALAYA SHELF, LOUISIANA, USA

By

Ilgar Safak

August 2010

Chair: Alexandru SheremetMajor: Coastal and Oceanographic Engineering

Interaction of near-bed wave-induced turbulence and cohesive sediments in muddy

environments is studied based on field observations and a bottom boundary layer

model. Wave, current, and sediment observations were collected with a suite of acoustic

and optical instrumentation at approximately 5-m depth on the muddy Atchafalaya

clinoform, Louisiana, USA. Low wave-bias estimates of near-bed Reynolds stresses are

obtained by a method that is based on differencing and filtering of velocities from two

sensors. The event that is focused on in this study is characterized by moderate waves

with high steepness, currents with speeds sometimes reaching 30 cm/s near bed, and

Reynolds stresses and suspended sediment concentrations reaching to their maximum

values throughout the experiment (0.4 Pa, 3 g/L). In general, Reynolds stresses are

found to be correlated with short-wave near-bed accelerations and suspended sediment

concentration, as previously observed on sandy beaches, where accelerations have

been associated with bed fluidization and sediment transport.

A detailed numerical analysis of the observations is performed with a one-dimensional

bottom boundary layer model for small scale turbulence and sediment transport

processes on cohesive beds. The model accounts for the coupling between the fluid

and the cohesive sediment phases, and uses a floc size that is constant in time and

space. A representative floc size is selected for the experiment site, based on two

independent sources that show consistency. Direct estimates of size distribution of

11

Page 12: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

suspended sediments in the vicinity of the experiment site show a remarkably stable

floc mode peak under varying wave and turbulence conditions. Indirect estimates of

equilibrium floc size are obtained through calculations of an analytical flocculation

model that uses observation-based parameters. With a floc size input based on the

observations, the model reproduces currents and suspended sediment concentrations

accurately; modeled Reynolds stresses match the low wave-bias estimates, with

better agreement for cases of stronger currents and smaller wave-orbital velocities.

The numerical simulations suggest that sediment-induced stratification effects are the

same order of magnitude as turbulent dissipation, and thus play a significant role in

the turbulent kinetic energy (TKE) balance within the tidal boundary layer. However,

inside the wave boundary layer, the ratio of stratification to shear-induced turbulence

production (i.e., gradient Richardson number) decreases significantly; shear-induced

turbulence production and turbulent dissipation dominate the TKE balance. For these

observations, model results show that the vertical structures of currents and Reynolds

stresses are relatively insensitive to the exact floc size.

Future efforts should include analysis of wider range of conditions (especially events

with higher near-bed concentrations), and comparison of model results with a more

detailed vertical structure of suspended sediment concentration.

12

Page 13: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

CHAPTER 1INTRODUCTION

1.1 Cohesive Sediments in Marine Environment

Marine sediment transport is a coastal process that affects shoreline change,

methods of coastal protection, design of coastal and offshore structures (see Sterling

and Strohbeck (1973) for an example of failure of oil platforms due to wave-induced

seafloor movements), underwater detection, navigation, water quality, and fate of

pollutants and biomatter due to settling of sediment from surface to deep sea (Hill,

1998). In the conventional scheme, bed sediment is entrained into the water column

by surface waves and is advected by currents. Recently, waves in shallow water were

also noted to cause a net sediment transport (e.g., Hoefel and Elgar, 2003; Hsu and

Hanes, 2004). According to the presence of cohesion between particles, sediments

are classified into two categories: cohesive, which is also commonly known as mud

(silt and clay with primary particle size smaller than 63 µm), and non-cohesive such

as sand grains and coarser particles (Figure 1-1). Mehta (2002) defines mud as

a mixture of water and sediment particles that are predominantly cohesive, which

exhibits a rheological behavior that is poroelastic or viscoelastic when the mixture is

particle-supported, and is highly viscous and non-Newtonian when it is in a fluid-like

state.

The majority of the coasts around the world are covered mostly with sand. However,

there are numerous muddy coasts dominated by cohesive sediments, especially

at river mouths, such as the Atchafalaya Shelf in the Gulf of Mexico (Allison et al.,

2000), continental shelf off the Eel River in Northern California (Traykovski et al.,

2000), southwest coast of India (Jiang and Mehta, 1996), east coast of China (Jiang

and Mehta, 2000), on the Amazon Delta (Cacchione et al., 1995), Po prodelta in the

Adriatic (Traykovski et al., 2007), etc. In muddy environments, there is strong evidence

of coupling between boundary layer turbulence and suspended sediment processes

13

Page 14: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

(Trowbridge and Kineke, 1994; Allison et al., 2000; Traykovski et al., 2000; Sheremet

and Stone, 2003; Sheremet et al., 2005; Allison et al., 2005; Kineke et al., 2006;

Traykovski et al., 2007; Jaramillo et al., 2009). Waves which generate shear stresses at

the bed greater than the bed strength cause liquefaction of a layer in the muddy bed with

a thickness depending on the properties of the bed material and the wave conditions

(Winterwerp et al., 2007). If the resulting cohesive sediment suspension near the bed

reaches a volume concentration of unity, a space-filling network with both fluid and solid

properties develops (Mehta, 1989). In the literature, cohesive sediment suspensions

with mass concentrations exceeding 10 g/L are commonly classified as fluid mud layers,

and this state is called structural density (Winterwerp and van Kesteren, 2004). The

internal friction within the fluid mud layer can dissipate the surface wave-generated

internal waves at the mud-water interface (Winterwerp et al., 2007). In the presence of

currents, fluid mud (formed and entrained by waves) is transported in the upper water

column, which otherwise remains just above the bed (Mehta, 1989). On the Eel River

continental shelf in Northern California (Traykovski et al., 2000; Wright et al., 2001),

on the Po prodelta in the Adriatic (Traykovski et al., 2007) and on the Yellow River

mouth in the Gulf of Bohai, China (Wright et al., 2001), wave induced fluid mud layers

of thickness on the order of the wave boundary layer thickness were observed to flow

downslope (offshore) in the form of gravity currents. Turbulence, necessary to prevent

deposition of the advected material by this mechanism, was generated by the flow

itself. On the shallow Atchafalaya Shelf, bed liquefaction and sediment resuspension by

surface gravity waves also result in the formation of fluid mud layers. These layers were

observed to flow in the form of a turbidity current (Jaramillo, 2008; Jaramillo et al., 2009).

Beyond the scope of this study, another mechanism through which mud affects

large scale nearshore ocean dynamics is the significant dissipation of wave energy

while propagating over muddy seafloors in shallow waters. Recent studies showed that

fluid mud layers dissipate wave energy not only within the swell band but also within

14

Page 15: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

the short-wave band. This is due to nonlinear energy transfers from high-frequency

bands towards low-frequency bands, which interact with the seafloor more significantly

(Sheremet and Stone, 2003; Sheremet et al., 2005; Kaihatu et al., 2007; Elgar and

Raubenheimer, 2008; Jaramillo, 2008; Sheremet et al., 2010). Figure 1-2 shows a

comparison of numerical simulation of wave field propagation on sandy and muddy

seafloors. Nonlinear interactions across spectrum are not accounted for. An idealized

case of unidirectional propagation over a distance of 5-km with constant 5-m depth

is set. An initial spectrum is selected based on the parameterizations obtained from

the directional wave measurements during the Joint North Sea Wave Project, i.e.,

JONSWAP (Hasselmann et al., 1980). A significant wave height of 2 m, and a peak

frequency of 0.1 Hz is selected (thick line in Figure 1-2). The material on the sandy

seafloor is represented by 0.2 mm grains. The muddy seafloor is set to have a 15-cm

thick viscous fluid mud layer, with a density of 1.1 g/cm3 and a kinematic viscosity of

10−3 m2/s (Kaihatu et al., 2007; Winterwerp et al., 2007). Bottom friction by the sandy

bed (Jonsson, 1966; Kamphuis, 1975; Dean and Dalrymple, 1991) causes visible

dissipation at only five frequencies around the spectral peak (thin line with squares in

Figure 1-2) and the resulting significant wave height is 1.8 m. The calculations, based

on a formulation which assumes the layer is a viscous fluid (Ng, 2000) show that the

muddy seafloor causes a more significant collapse in the wave field’s energy, in a wide

range of frequencies (dashed line with dots in Figure 1-2). The resulting significant

wave height for this case is 0.6 m. Therefore, while bottom friction by the sandy bed

dissipates 19% of the initial energy of the wave field, the muddy seafloor, characterized

by the parameters given above, dissipates 91% of the initial energy.

In terms of small scale ocean bottom boundary layer processes, interaction of

turbulent flow with sediment is more challenging to study in muddy environments than in

sandy environments because of the complicated physics related to cohesive sediments.

In a cohesive sediment suspension, rather than bouncing away from each other as

15

Page 16: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

sand grains do, the collision of sediment particles results in the formation of aggregates

called flocs. These flocs are characterized by high water content (Winterwerp and

van Kesteren, 2004) and a fractal geometry with dimension close to 2 (Kranenburg,

1994). The flocculation process is discussed in more detail in the following sections.

The settling velocity of sand grains can be calculated as a function of grain size and

fluid viscosity through Stokes’ law (e.g., Nielsen, 1992). However, changes in the

water content, geometry and, therefore, density of cohesive sediment flocs cause

their settling velocity to be temporally varying, and higher than those obtained for

primary particles, by Stokes’ law. Therefore, a cohesive sediment sample needs to

be considered site-specific and event-specific in the sense that flocs in each sample

may have varying physical properties such as size, density, and settling velocity

(Mehta, 1989). Gases and organic particles in cohesive sediments (Winterwerp and

van Kesteren, 2004) should also be considered in the analysis of cohesive sediments

interacting with fluid flow. In fact, the interest of the U.S. Navy in the hydrodynamics in

muddy environments has recently increased due to the fact that available Navy models

for waves and circulation were mostly developed for sandy environments but do not work

well with mud.

1.2 Turbulence in Combined Wave-Current Flow

Flow-generated turbulence is a major mechanism in large scale processes such

as momentum balance in the surf zone (e.g., Trowbridge and Elgar, 2001) and small

scale sediment transport processes (Winterwerp, 1998). Numerical models for ocean

bottom boundary layer turbulence are often based on two-equation turbulence closures

and validated with laboratory experiments (e.g., Winterwerp, 2001; Hsu et al., 2007,

2009). On shallow continental shelves, surface wave-induced orbital velocities at the

bed generate a wave boundary layer of a few centimeters thick, much thinner than

the current boundary layer that scales with water depth (e.g., Fredsoe and Deigaard,

1992). Waves cause currents to experience a bottom drag above the wave boundary

16

Page 17: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

layer larger than that associated with the bottom roughness, deviating from the

logarithmic vertical structure and having a reduced vertical shear near the bed (e.g.,

Grant and Madsen, 1979, 1986). The resulting non-linear friction processes and the

differences in the characteristic length and time scales of waves and currents complicate

modeling of turbulence and sediment transport processes in the bottom boundary layers

(e.g., Styles, 1998; Styles and Glenn, 2000, 2002; Hsu et al., 2009). This is a major

research area in oceanographic studies because sediment entrainment from the bed

is controlled by wave boundary layers where a significant amount of energy is being

dissipated (Trowbridge and Agrawal, 1995) and exchanges of heat and momentum

occur. The flow parameters in these boundary layer models (e.g., eddy viscosity, see

Winterwerp (2001) and references therein) are modified by the turbulence-damping

effect of density-induced stratification, i.e., vertical gradient of suspended sediment

concentration. The turbulence closures on which these models are built have yet to be

evaluated in detail against high-resolution field observations of combined wave-current

flows and sediment transport processes.

1.3 Interaction of Turbulent Flow with Cohesive Sediments

Hydrodynamic properties of cohesive sediments, which differ significantly from

those of the primary particles (e.g., settling velocity, Section 1.1), result in stratification

effects on vertical mixing that are specific to cohesive sediments (e.g., hindered

settling, formation of near-bed fluid mud layers). In hypothetical cases of cohesive

sediment suspensions in tidal flow, Winterwerp (2001) numerically simulated the

formation of these fluid mud layers when the concentration of suspension exceeded

the sediment-carrying capacity of the flow. Simulations in that study showed that

stratification modifies vertical structures of velocity and turbulent flow parameters at

depth-averaged concentrations as low as 0.1 g/L, as long as near-bed fluid mud layers

of at least a few centimeters thick are formed. The one-dimensional boundary layer

model, used by Winterwerp (2001), was validated with data from laboratory experiments

17

Page 18: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

of cohesive sediment suspensions of depth averaged concentrations of 1 g/L in tidal

flow (Winterwerp, 2006). Model calculations of Winterwerp (2006) agreed with the

measured velocity and suspended sediment concentration; the model also captured

the decreasing trend of vertical eddy viscosity (estimated through measured data) with

increasing sediment concentration. In other related studies on steady currents in the

tidal boundary layer, field observations (e.g., Trowbridge and Kineke, 1994), laboratory

experiments (e.g., Gratiot et al., 2005), and models (e.g., Trowbridge and Kineke, 1994;

Michallet and Mory, 2004) seem to agree that turbulence is significantly affected by

cohesive sediment in the vicinity of a steep vertical gradient of suspended sediment

concentration. These gradients are commonly known as a lutocline, which separates a

high concentration layer with a mixed and comparatively low concentration layer (e.g.,

Parker and Kirby, 1982; Mehta, 1989; Vinzon and Mehta, 1998).

Flocculation is a combined process of aggregation (floc formation due to collision

of particles by turbulent motions) and floc breakup (disruption of flocs by turbulent

shear), which is governed by turbulent eddies scaled by the Kolmogorov microscale

(Berhane et al., 1997). While the importance of flocculation in altering the flow-sediment

interaction through stratification is well established (Winterwerp, 2001), the role of the

floc size (and therefore settling velocity of the floc) is less understood. Dyer (1989)

hypothesized that floc structure is a function of turbulence levels and availability of

sediment, i.e., suspended sediment concentration. The argument was that flocs should

grow as the concentration of primary particles increases, due to increased probability of

particle collisions. Turbulence (low to moderate shear stresses) is needed to promote

collisions; however, high turbulence levels (high shear stresses) are expected to break

the flocs and limit their size. In turn, floc structure should affect the hydrodynamic

properties of flocs (e.g., settling velocity) with direct effects on the residual time of the

flocs in the water column and, therefore, the vertical structure of suspended sediment

concentration. Following the hypothesis that increasing turbulent shear stress first

18

Page 19: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

increases and then decreases floc size (Dyer, 1989), simplified analytical expressions

for equilibrium values of floc size and settling velocity were proposed by Winterwerp

(1998), for a constant fractal dimension of 2. The formulation of Winterwerp (1998)

was later modified by Son and Hsu (2009) to include variable fractal dimension and

variable floc yield strength. The simplified formulation using the floc fractal dimension

2 was used to model cohesive sediment transport in the boundary layer of a tidal

channel (Winterwerp, 2002). Winterwerp (2002), however, did not discuss the details

of the turbulence kinetic energy balance. The flocculation models discussed above

were calibrated for cohesive sediment transport in the laboratory with simple shear

flow or homogeneous turbulence and applied to a tidal boundary layer condition. Their

applicability to wave-induced fluid mud transport is unclear. Moreover, turbulence may

not be the governing mechanism that controls flocculation at every condition. Based

on field observations, Hill (1998) proposed a flocculation mechanism with a different

dependence on turbulence: at low turbulent energy, disruptive stresses on flocs due to

sinking in the fluid may exceed turbulence-induced stresses and limit floc size.

1.4 Objectives of This Study

In this study, the interaction between turbulence induced by combined wave-current

flow and suspended cohesive sediments in bottom boundary layers of muddy environments

is studied based on observations collected during spring of 2006 and 2008, on the

Atchafalaya inner shelf, Louisiana, USA. Wave and current parameters are calculated

using standard data assimilation methods and spectral analysis. In addition, near-bed

Reynolds stresses are estimated through the data with the most advanced of a series of

methods recently proposed to reduce wave bias in turbulence estimates (Feddersen and

Williams, 2007). This is a challenging task, especially in wave-energetic environments

(Trowbridge, 1998). Having Reynolds stress estimates provides the advantage of

designating, merely by analyzing the observations, the events when turbulent fluxes in

the boundary layer are likely to be significant. Based on these observations, small-scale

19

Page 20: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

turbulent flow and sediment transport processes are modeled using a one-dimensional

bottom boundary layer model for cohesive beds (Hsu et al., 2007, 2009). The model

uses a constant floc size, and therefore a constant settling velocity, in contrast with other

models (e.g., Winterwerp, 2002) that predict a settling velocity varying over a tidal cycle.

While this may lead to errors when used to simulate the vertical structure of suspended

sediment concentration, the constant floc size assumption seems to be consistent

with the field observations, which suggest a weak relationship between turbulence

and floc size, at least in relatively dilute suspensions. Direct observations of floc size

distribution (but not co-located with these particular hydrodynamic measurements)

are used to estimate a representative floc size. The values used here are of the same

order of magnitude as the equilibrium floc sizes calculated with the model proposed

by Winterwerp (1998) using the observed primary particle size, suspended sediment

concentration and estimated Reynolds stresses.

The description of the field experiment, the details of instrumentation and sampling

schemes, an overview of the sedimentological characteristics of the experiment site,

typical atmospheric and flow conditions for the site, and a brief summary of the wave

and near-bed flow observations during the entire experiment period are presented in

Section 2. In Section 3, the data analysis methods used are presented. A general

overview of the observations during the experiment studied herein is given in Section

4, together with the details of a 1-day event which is characterized by the highest

amount of suspended sediment recorded near bed. Why a relatively stronger local

turbulence-cohesive sediment interaction is expected during this event is discussed.

The bottom boundary layer model is presented in Section 5, together with the

governing equations, description of its execution procedures, and discussions on

the selection of the floc size input based on two independent sources. Capabilities

of the model are tested first by comparing the model results with the measurements.

Model calculations of Reynolds stress are compared with the observation-based

20

Page 21: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

estimates. The contribution of different terms in the turbulence kinetic energy balance

and sensitivity of this balance on floc size are evaluated. The results are summarized

and discussed in Section 6.

21

Page 22: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

Figure 1-1. Noncohesive sand grains and cohesive fluid mud. (a) Sand grains; and (b)mud sample collected at the Atchafalaya Shelf, Louisiana (Photo courtesy:K. T. Holland, Naval Research Laboratory at the Stennis Space Center).

22

Page 23: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

0 0.1 0.2 0.3 0.4 0.5 0.610

−2

10−1

100

101

frequency (Hz)

flux

spec

tral

dens

ity (

m3 )

Figure 1-2. Comparison of spectral evolution of waves over muddy and sandy seafloors.Initial JONSWAP spectrum (thick continuous line), resulting spectra afterpropagating 5-km in 5-m depth over a flat bottom covered with 0.2 mm sand(thin continuous line with squares), and 15-cm thick viscous fluid mud layer(dashed line with dots).

23

Page 24: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

CHAPTER 2FIELD EXPERIMENT

2.1 Experiment Site

The main data set on which this study is based was collected from March 25th to

April 7th, 2008. This was the last two-week interval of an experiment which started on

February 22nd, 2008 on the muddy inner shelf fronting Atchafalaya Bay, Louisiana,

USA, in the north-central Gulf of Mexico (Figures 2-1 and 2-2). This experiment was

part of a larger scope study of wave, turbulence and sediment transport processes in

shallow muddy environments that started in 2006. Four instrumented platforms were

deployed on February 22nd by the University of Florida and the University of Texas

(Figure 2-2). Platforms 1-3 were put in a cross-shore transect and Platform 4 was

located in an alongshore transect with the shallow platform 3. Platform 5 was near

Fresh Water Bayou about 60 miles west of the Atchafalaya Bay, near the center of a

transect of current-meters, deployed by the group led by Dr. Steve Elgar and Dr. Britt

Raubenheimer from Woods Hole Oceanographic Institution (Safak et al., 2010b). The

coordinates and retrieval dates of the platforms are given in Table 2-1, together with the

mean depth at the location of the platforms (average of the data collected throughout the

experiment). The experiment site was revisited for each two-week period to retrieve data

and change the batteries of the instruments deployed at platforms 1-4. Therefore, the

three periods February 22nd - March 8th, March 8th - March 25th and March 25th -April

7th were named as experiments A,B and C, respectively. The focus of this study is the

observations collected at Platform 2 near the 5-m isobath during Experiment C.

The experiment site is located on the topset of the Atchafalaya sub-aqueous

feature, which is defined as a clinoform of up to 3-m thick mud layer. The clinoform

extends out to the 8-m isobath at tens of kilometers offshore (Allison et al., 2000; Neill

and Allison, 2005). Since the 1940s, the muddy Atchafalaya inner continental shelf

receives about 30% of the discharge of the Mississippi River, which is the largest river

24

Page 25: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

on the North American continent (Mossa, 1996; Allison et al., 2000; Neill and Allison,

2005). The Atchafalaya River leaves the main course 320 km upstream of the Gulf of

Mexico near Simmesport, Louisiana. The annual sediment discharge by the Atchafalaya

River is estimated to be 84 million metric tons. The suspended sediment load into the

bay and the inner shelf is identified by fine grains with median particle diameter D50=2-6

µm, however, includes 17% sand, as well (Allison et al., 2000). West of Marsh Island

(92.5 degrees longitude West), D50 was estimated to be between 2.8-5.9 µm (Allison

et al., 2005). Sheremet et al. (2005) measured D50=6.34 µm on the inner shelf, about

20 km offshore of Marsh Island. These data suggest that D50=2-7 µm is representative

for the area. Size distribution of suspended sediments is available from the previous

experiments and is investigated in the following sections.

The Atchafalaya River is different than the major distributary of the Mississippi 180

km to the east in the sense that the Atchafalaya Shelf is shallower with a milder slope

(e.g., low-gradient where the 10-m isobath is approximately 40 km offshore) and more

wave-energetic. This region is interesting for sediment transport studies due to the

fact that over the last few decades, the coastline has been prograding seaward with

rates of O(10 m/yr) and land accretes vertically at rates reaching O(1 cm/yr) due to

the sediment discharge of the Atchafalaya River in spite of rising sea levels. However,

much of the rest of the Louisiana coastline and the Mississippi Delta are experiencing

significant erosion (Allison et al., 2000; Draut et al., 2005; Neill and Allison, 2005; Kineke

et al., 2006). Despite these high sediment accumulation rates, it is unlikely that the

outer Atchafalaya Bay and the clinoform topset area will accrete to sea level, since

atmospheric conditions force the hydrodynamics to distribute sediments away from

these areas (Neill and Allison, 2005).

Wind is the major forcing that controls circulation, sediment transport, water

level, and salinity changes over the Atchafalaya Shelf. The wave climate is dominated

between December and April (coinciding with the period of high sediment discharge in

25

Page 26: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

the Atchafalaya River) by wave fields associated with storms and cold fronts passing

through the area on 3-7 day time scales (Allison et al., 2000; Walker and Hammack,

2000). These fronts are characterized by pre-frontal onshore winds, strong wave

activity and coastal setup, followed by post-frontal offshore winds, set-down, and

formation of large sediment plumes (Allison et al., 2000; Walker and Hammack, 2000).

At the experiment site, these perturbations can generate swells in excess of 1-m

height lasting for several days. In shallow water, such intense wave activity causes

significant variations of the bed state throughout a storm, in a sequence of breaking

down stratification, triggering bed liquefaction, increasing sediment resuspension and

turbidity throughout the water column (Allison et al., 2000), followed by the formation

of fluid mud layers within 1-m near the bed due to settling, and finally consolidation to

a soft bed (Jaramillo et al., 2009). These high concentration fluid mud layers can be

transported over the shelf in different directions by various mechanisms: westward by

residual currents of about 10 cm/s which may account for advection of more than half

of the sediment discharge into the shelf (Wells and Kemp, 1981; Allison et al., 2000);

onshore by coastal upwelling (Kineke et al., 2006); and offshore in the form of a turbidity

current of about 5 cm/s, which is maintained in suspension by wave-induced turbulence

(Jaramillo et al., 2009). On the other hand, the contribution of extreme events such as

tropical cyclones and hurricanes (e.g., Hurricane Lili in 2002) to sediment transport and

organic matter deposition may exceed the input by the Atchafalaya River (Allison et al.,

2005; Goni et al., 2006).

Together with the studies on nearshore circulation and sediment transport

processes discussed above, the broad-spectrum wave-energy dissipation effect of

muddy seafloors has been studied extensively on the Atchafalaya Shelf and near Fresh

Water Bayou in the last decade (Sheremet and Stone, 2003; Sheremet et al., 2005;

Elgar and Raubenheimer, 2008; Jaramillo, 2008). A recent finding from these wave-mud

interaction studies is that maximum mud-induced dissipation rates are observed not

26

Page 27: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

during the input of highest wave energy into the system but in the wane of the storm.

During that period, the seafloor is characterized either as a viscous fluid or a viscoelastic

material, and the boundary layer is expected to be laminar (Jaramillo, 2008; Sheremet

et al., 2010).

2.2 Instrumentation

A typical platform in 2006 and 2008 experiments is shown in Figure 2-3. Currents

in the upper water column and the directional surface wave field were monitored by

upward-pointing Acoustic Doppler Current Profilers (ADCP, Teledyne RD Instruments,

1200 kHz, Figure 2-3 label A). Velocity and backscatter within 1-m of the bed were

continuously measured by downward-pointing Pulse Coherent-Acoustic Doppler

Profilers (PC-ADP, Sontek/YSI, 1500 kHz, Figure 2-3 label B) in bins of either 1.6- or

3.2-cm continuously with sampling rates of either 1- or 2-Hz. The PC-ADPs are also

equipped with built-in pressure sensors. Acoustic Backscatter Profilers (ABS, Aquatec,

Figure 2-3 label C) provided a more detailed profile of the near-bed backscatter at

bins smaller than 1 cm. Optical backscatterance sensors (OBS-3s and OBS-5s, D &

A Instruments, Figure 2-3 label D) measured turbidity. Operational principles of these

acoustic and optical instruments can be found in Lhermitte and Serafin (1984) and

Downing et al. (1981), respectively. Conductivity-Temperature sensors (CT, SeaBird

Electronics) provided salinity estimates and temperature measurements. At the location

of Platform 1, a pressure sensor, located 1-2 m below the surface, sampled at 4-Hz

throughout the entire 7-week experiment. At the location of Platform 3, an Onset, Inc.,

HOBO micro-station located at an elevation of 7-m above the sea surface, provided

30-min averages of wind speed and direction.

The instrumentation used in this analysis comprised a vertical array of two

synchronized Acoustic Doppler Velocimeters (ADV, SonTek/YSI Hydra 5-MHz), and

an OBS-5. An ADV is a pointwise velocity sensor, and it has been tested in laboratories

to estimate turbulence parameters (Voulgaris and Trowbridge, 1998) and to measure

27

Page 28: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

instantaneous velocities in concentrated fluid mud (Gratiot et al., 2000). A photograph

of the instrumentation on the platform before the deployment and a schematic of the

platform showing the locations of the instruments are shown in Figures 2-4 and 2-5,

respectively. The sampling volume of the lower ADV (ADV-1) was located at 17 cmab

(cm above bed); the higher ADV (ADV-2) sampled in a volume at 145 cmab. Based

on the measurements in a wave-free environment (Trowbridge et al., 1999), Shaw

and Trowbridge (2001) suggested that two vertically-stacked sensors are optimally

configured to yield uncorrelated cross-sensor turbulent covariances if the distance

between the sensors is greater than 5 times the height of the lower sensor above

the bed. For the application of the two-sensor method of wave-bias reduction from

the Reynolds stress estimates (see Section 3.2.3), the configuration of the ADVs

(Figure 2-5) follows this recommendation. Each ADV was equipped with a built-in

pressure sensor, located at about 60 cm from its sampling volume. The ADVs sampled

pressure and three-dimensional flow velocity (converted to East-North-Up coordinates

in post-processing) at 10 Hz, in 10-min measurement bursts, one burst every hour, for

the entire two-week duration of the experiment. A burst duration of 10-min allows the

ADVs to run for two weeks with a shared battery pack, spans about 75 swell periods

(for a swell period of 8 s, which is typical for the site), and is long enough to provide a

stable mean current estimate (Soulsby, 1980). The OBS-5 was mounted at 12 cmab,

and recorded 1-min averages of backscatter at a rate of 2 Hz. The backscatter signal

from the OBS-5 was calibrated in the laboratory with in-situ sediment and water samples

collected at the site surveys during the deployment and retrieval of the platforms.

Through this calibration, the turbidity records were converted to suspended sediment

concentration estimates. For the directional wave measurements, the data collected

by the ADCP at Platform 3, located 4 km onshore of the experiment site (Figure 2-2)

was used. The ADCP transducer head was located at 1.3 mab and measured pressure,

acoustic surface track, and velocity profiles at 2 Hz, in 40-min measurement bursts, one

28

Page 29: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

burst every hour. Wave data were processed using the Teledyne RDI software packages

WavesMon and WaveView, with a frequency resolution of 0.0078 Hz and an angular

resolution of 4 degrees. The ADCP also provided 10-min averaged current profiles in

bins of 20 cm.

For this analysis, the only information about the size distribution of suspended

sediments on the Atchafalaya Shelf was available from a LISST-100X Type-C (Laser In

Situ Scattering Transmissometer, Sequoia Scientific). LISST records the small-angle

scattering distribution of particles in water, which is inverted into size spectra. For further

operational principles, see Agrawal and Pottsmith (1994). Two sets of observations

(Allison et al., 2010) were collected, one from the 2006 experiment between February

28th and March 14th (Jaramillo et al., 2009) at the location of Platform 3 (Figure 2-2)

and one from the 2008 experiment A between February 22nd and March 8th, at Platform

4 (Figure 2-2). As both data sets have time offsets with the analyzed experiment, the

2006 data set was preferred because it was collected at a location closer to the location

of the data set used in the analysis herein. The LISST background was calibrated using

filtered water at the deployment site. The path of the LISST measurements (which

determines the upper limit of sediment concentration at which reliable data can be

obtained) is reduced with an 80% path reduction module. With this setting, the threshold

sediment concentration is about 1 g/L, and reliable data were collected at 120 cmab

between March 1st and March 9th, 2006. The instrument estimates size distributions

of suspended particulates (flocs and primary) in 32 class ranges between 2.5-500

µm size. However, in this experiment, the data is unreliable above 350 µm, and not

reproduced here. The grain-size distribution was recorded every minute (an average

of 100 samples at 2 Hz) in 30-min bursts each hour. On the same platform with the

LISST, a downward-pointing PC-ADP measured velocity and backscatter profiles,

and pressure. The PC-ADP sampled at 2 Hz sampling in 60 bins of 1.6 cm, following

a 10-cm blanking distance. A 10-min burst was started every 30-min. The PC-ADP

29

Page 30: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

measurements were missing velocity profiles once in every three or four samples, due to

an unknown instrumentation problem. Therefore, the 2 Hz velocity measurements were

burst-averaged to calculate the vertical structure of mean currents and then estimate

the bottom friction velocity by using the logarithmic law of the wall (Section 3.3). These

estimates, and the spectral wave calculations based on the PC-ADP pressure sensor

data, present a general picture of the variation of size distribution records of LISST

under varying wave and bottom turbulence conditions (Sections 3.3 and 4.2).

2.3 General Conditions

Before focusing on the experiment of interest and the presentation of the related

data analysis methods, a general overview of the conditions (winds, near-bed flows,

and surface waves) throughout the entire seven-week experiment duration is presented

in Figure 2-6. For this, the wind and the PC-ADP data measured at Platform 3 (Figure

2-2), where the local depth was about 3.8 m, and the pressure data from the pressure

sensor near Platform 1 (Figure 2-2), where the local depth was about 7.4 m, are used.

The PC-ADP on Platform 3 was the only near-bed profiler that sampled throughout the

entire experiment, except the retrievals for re-deployments between experiments A-B

and B-C. The pressure sensor near Platform 1 was the only wave gauge which sampled

continuously throughout the entire experiment. In the wave spectral calculations, which

are detailed in the next section, swell (low frequency) and sea (high frequency) bands

are decomposed using a cutoff frequency of fc=0.2 Hz. This arbitrary value is selected

based on the similar evolution trends of spectral wave energy at frequencies larger

than this value, and wind speed. Figures 2-6a and b show that sea waves (black curve

in Figure 2-6b) closely follow the trend of wind speed, which sometimes exceeded 15

m/s and generated sea waves of significant heights exceeding 1.5 m (see the peaks

on February 28th, March 4th, March 8th, and March 20th). Increasing swell energy

seems to be triggered by shifts in wind direction from onshore winds to offshore winds

on February 27th, March 4th, March 8th, and especially March 18th (see the variation

30

Page 31: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

in the wind-direction representing color code in Figure 2-6a, from cyan-green sector,

i.e., northward winds, to red-magenta sector, i.e., southward winds). During all four

of these events, near-bed activity was observed in the sense that the location of the

maximum backscatter is recorded to be closer to the PC-ADP sensor head compared

to its initial location at the beginning of the experiment (Figure 2-6c). This indicates

the formation of a high concentration, i.e., fluid mud layer above the consolidated bed.

The backscatter is also showing an increase througout the water column during these

events. The strongest swell event throughout the experiment was recorded on March

18th when swells reached 1.5 m and the near-bed observations indicate formation of

a fluid mud layer of about 20 cm thickness. During the third and last 2-week interval of

this experiment, experiment C, the data set for studying turbulence-cohesive sediment

interaction herein was collected. During that experiment, swell heights were mostly less

than 0.5 m at 7.4 m depth and the backscatter records are not indicating formation of

any fluid-mud layers at 3.8 m depth.

31

Page 32: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

Table 2-1. Location, mean depth, and retrieval dates of the instrumented platformsPlatform Latitude(North) Longitude (West) Depth (m) Retrieval date

1 29o11.815’ 91o36.731’ 7.4 March 25th2 29o13.439’ 91o34.807’ 5.0 April 10th3 29o15.574’ 91o34.267’ 3.8 April 10th4 29o22.238’ 91o46.922’ 4.0 April 10th5 29o33.759’ 92o33.800’ 4.0 March 25th

32

Page 33: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

−95 −90 −85 −8025

26

27

28

29

30

31

Longitude (deg.)

Latit

ude

(deg

.)

Florida

Atlantic Ocean

Louisiana

Gulf ofMexico

Transect 1Transect 2

Figure 2-1. Gulf of Mexico. The platforms along Transect 1 and Transect 2 werecontaining instrumentation deployed by the field support groups of theUniversity of Florida-the University of Texas, and the University ofFlorida-Woods Hole Oceanographic Institution, respectively.

33

Page 34: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

Longitude (deg.)

Latit

ude

(deg

.)

Marsh Island

Atchafalaya River

Louisiana

Atchafalaya Bay

Trinity shoal

Gulf ofMexico

123

4

5

5 m10 m

20 m

30 m

−92.8 −92.4 −92.0 −91.6 −91.228.8

29.0

29.2

29.4

29.6

29.8

30.0

Figure 2-2. Bathymetry of the Atchafalaya Shelf and the locations of the instrumentedplatforms. This bathymetry was based on a data set collected before the2008 experiment and is showing differences with the 2008 bathymetry. Thenumbered red stars indicate the platform locations. The main data set onwhich this study is based was collected at Platform 2. Also from Platform 3which is located 4 km onshore of Platform 2, directional wavemeasurements, wind data, and information on size distribution of suspendedsediments in the area are investigated.

34

Page 35: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

Figure 2-3. A typical instrumentation platform. The platforms were equipped with anupward-pointing ADCP (A), a downward-pointing PC-ADP (B), adownward-pointing ABS (C), and pointwise turbidity sensors (D).

35

Page 36: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

Figure 2-4. The synchronized ADVs and the OBS-5 on the platform prior to thedeployment.

36

Page 37: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

Figure 2-5. Configuration of the instrument array. Circles mark the location of thesampling volumes.

37

Page 38: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

Figure 2-6. Wind, wave, and near-bed conditions on the Atchafalaya Shelf throughoutthe seven-week experiment in Spring 2008. (a) Wind speed and direction(the color code shows where the wind flow is towards) measured nearPlatform 3 where the local depth is 3.8 m; (b) significant wave height in sea(black) and swell (red) bands, measured by the near-surface pressuresensor at Platform 1 where the local depth is 7.4 m; and (c) near-bedacoustic backscatter (normalized such that the maximum backscatter isshown by dark red and the minimum backscatter by dark blue) measured bythe PC-ADP at Platform 3. The gaps in the backscatter data on March 8thand March 25th correspond to the intervals between the experiments A-Band B-C, when the data sets from the instruments were being retrieved, andthe batteries and memory cards were being replaced.

38

Page 39: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

CHAPTER 3DATA ANALYSIS METHODS

3.1 Wave Spectral Calculations

A wave mode with amplitude a and frequency f has a total average energy (sum of

kinetic and potential energies) per unit surface area (Dean and Dalrymple, 1991):

E (f ) =ρg a2

2, (3–1)

where ρ is the fluid density and g is the acceleration caused by gravity. In the frequency

domain, wave energy is usually represented as:

S(f ) =a2

2 df, (3–2)

where S(f ) is the energy spectral density of the mode with frequency f , and df is

the frequency resolution. The wave field during the analyzed experiment is defined

by statistical parameters which are obtained through spectral calculations. Also, the

cross-spectrum of horizontal and vertical velocity measurements is calculated to

estimate Reynolds stresses (Section 3.2). Therefore, a brief discussion of spectral

analysis is given here. Spectral analysis is used to decompose a time-varying quantity

x(t) into a sum of sine and cosine functions, and calculate the distribution of energy

at modes of different frequencies (Priestley, 1981). For a stochastic process, the

auto-correlation function describes the general dependence of the data values at one

time on the values at another time (Bendat and Piersol, 1971). It is defined as:

Rxx (τ) =∞∫

−∞

x(t) x∗(t + τ)dt , (3–3)

where t is the time and asterisk denotes the complex conjugate. Fourier transform of

x(t) at frequency f is:

39

Page 40: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

X (f ) =∞∫

−∞

x(t) e−2π�tdt , (3–4)

where i is the imaginary number. The Fourier transform of the auto-correlation function

gives the energy spectral density:

Sxx (f ) =∞∫

−∞

Rxx (τ) e−2π�tdt = 〈X (f )〉2 , (3–5)

where 〈〉 denotes the expected value operator, i.e., mean. While the auto-correlation

function describes the correlation within a process (equation (3–3)), the cross-correlation

function describes the correlation structure between two stochastic processes (Priestley,

1981), say x(t) and y (t):

Rxy (τ) =∞∫

−∞

x(t) y ∗(t + τ)dt. (3–6)

The Fourier transform of the cross-correlation function gives the cross-spectral density:

Cxy (f ) =∞∫

−∞

Rxy (τ) e−2π�tdt = X (f )Y ∗(f ) , (3–7)

which is decomposed into its real part, i.e., co-spectrum, and imaginary part, i.e.,

quadrature spectrum, as follows:

Cxy (f ) = Coxy (f ) + i Quadxy (f ). (3–8)

Spectral density of pressure (Spp) is estimated using standard discrete Fourier

spectral analysis based on the pressure time series from ADV-2. The pressure time

series from each 10-min burst is demeaned and detrended, divided into 51.2-sec

segments (each containing 1024 samples) with 50% overlap. The purpose of overlapping

is to reduce leakage at the boundaries of the segments. The resulting spectral estimates

are characterized by approximately 22 degrees of freedom and 0.0196 Hz frequency

40

Page 41: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

resolution. Spectral density of the free surface elevation (Sηη) is estimated from Spp by

correcting for depth attenuation, using linear wave theory:

Sηη(f ) =[

cosh(kw h)cosh kw (h + zsensor )

]2

Spp(f ) , (3–9)

where kw is the wavenumber, h is the local water depth, and zsensor is the elevation of

the sensor relative to the mean water level. The wavenumber is related to the frequency

through the linear dispersion relation for waves:

(2πf )2 = gkw tanh (kw h) . (3–10)

A cutoff frequency is defined by a variance attenuation with depth of less than 2% from

surface to sensor, then a spectral tail proportional to f −5 (Phillips, 1958) is added to

cover the high-frequency range. The band significant wave height (Hs) is estimated

using the standard relation:

Hs = 4

√√√√√f2∫

f1

Sηη(f ) df , (3–11)

where f1 and f2 are the spectral band limits. Velocity and acceleration spectral densities

are estimated with the same procedure but from the measurements of ADV-1, as this

sensor was placed closer to the bottom. Velocity and acceleration variance estimates

are calculated as:

σ2ξ =

f2∫

f1

Sξξ(f ) df , (3–12)

where ξ represents the parameter of interest, i.e., velocity or acceleration component.

Where convenient, swell (long waves) and sea (short waves) bands are decomposed by

using a cutoff frequency of fc=0.2 Hz, such that the swell band is defined by f1=0.0196

Hz, f2 = fc .

41

Page 42: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

3.2 Estimation of Reynolds Stresses in the Presence of Surface Waves

3.2.1 Definition of Reynolds stress

In the Reynolds-averaged Navier-Stokes equations, the contribution of the turbulent

motion to the mean stress is described with the components of the symmetric Reynolds

stress tensor τij (also referred to as turbulent covariance tensor), which includes

correlations of turbulent fluctuations (Tennekes and Lumley, 1972):

τij = −ρ

u′u′ u′v ′ u′w ′

v ′u′ v ′v ′ v ′w ′

w ′u′ w ′v ′ w ′w ′

, (3–13)

where u and v are the horizontal components and w is the vertical component of

velocity, a prime denotes turbulent fluctuations, and an overbar denotes a time-averaged

quantity. The diagonal components are normal stresses, i.e., pressures. These normal

stresses contribute little to the mean momentum transport (Tennekes and Lumley,

1972), however, half of the sum of these stresses is defined as the turbulent kinetic

energy (k), which is one of the most common quantities used in two-equation turbulent

flow modeling (Section 5):

k =12

(u′2 + v ′2 + w ′2

). (3–14)

The off-diagonal components of equation (3–13) are Reynolds shear stresses, which

play a dominant role in momentum transport in turbulent motion (Tennekes and Lumley,

1972). For the one-dimensional bottom boundary layer modeling study herein, the

Reynolds shear stresses of interest are those that appear in the horizontal momentum

equations:

τxz = −ρu′w ′ ; τyz = −ρv ′w ′ . (3–15)

42

Page 43: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

Due to their significant effects in boundary layer flows at several scales such as

large-scale surf zone flows (Trowbridge and Elgar, 2001) and small-scale sediment

transport processes (Winterwerp, 1998), estimating Reynolds stresses accurately using

field observations in the absence of models (uniform eddy viscosity model, mixing length

model, two-equation models, etc.) is an important objective in oceanographic studies. In

addition, as mentioned in Section 1.3, the closure schemes implemented in turbulence

models to calculate the Reynolds stresses need to be evaluated against observations of

combined wave-current flows.

3.2.2 Wave bias in Reynolds stress estimates

The main difficulty in estimating Reynolds stresses to characterize flow turbulence

is due to the contamination of the turbulent signal by surface waves, which typically

dominate the variance of horizontal and vertical velocities (Trowbridge, 1998). The

discussions in the rest of this section are going to be for a two-dimensional model

velocity vector u = (u, w ). Therefore, the quantity of interest is u′w ′ (equation (3–15)),

however, the analysis is easily adaptable to v ′w ′ as well. u is decomposed as:

u = �u + ~u + u′ , (3–16)

where an overbar denotes a time-averaged quantity, i.e., mean current, a tilde denotes

wave-induced fluctuations, and a prime denotes turbulent fluctuations, respectively.

Assuming that wave-induced and turbulence-induced fluctuations are uncorrelated, i.e.,

statistically orthogonal (Kitaigorodskii and Lumley, 1983), the quantity of interest u′w ′

can be obtained as:

u′w ′ = uw − uw . (3–17)

The first term on the right hand side is the covariance of the measured horizontal

and vertical velocities. However, uw needs to be separated in the presence of energetic

wave motions. Several approaches have been developed to reduce the wave bias in

43

Page 44: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

the Reynolds stress estimates. Ensemble averaging is implemented in the time domain

and wave-induced fluctuations are estimated by averaging the same point in the wave

phase over several consecutive regular waves. High-pass filtering can be applied by

assuming that waves and turbulence scales are clearly separated, and then specifying

a cutoff frequency. The accuracy of this approach depends on how carefully the cutoff

frequency is selected, since effects of large-scale turbulent eddies may be neglected.

These two techniques can be used in tightly controlled environments such as laboratory

experiments (Scott et al., 2005), but are less useful for random wave fields. In a class

of methods, the wave signal is identified by velocity fluctuations that are coherent with

pressure, whereas, fluctuations that are not coherent with pressure are considered to

be due to turbulence (Benilov and Filyushkin, 1970; Agrawal and Aubrey, 1992; Wolf,

1999). The coherence between the pressure and the wave-induced velocity fluctuations

is calculated as:

γ2 =CpuC ∗

pu

SppSuu, (3–18)

where Cpu is the cross-spectrum of the pressure and velocity (equations (3–7) and

(3–8)), Spp and Suu are power spectra of pressure and velocity, respectively. The power

spectrum of turbulence-induced velocity fluctuations can be calculated as (Agrawal and

Aubrey, 1992):

Su′u′ = Suu[1− γ2] . (3–19)

The only additional requirement of this method is to have pressure measurements

synchronized with velocity measurements. However, this approach requires accurately

resolving wave nonlinearities and directional spread (Herbers and Guza, 1993).

Therefore, its practical applicability for field studies is questionable. The Phase method

(Bricker and Monismith, 2007), on the other hand, requires a single pointwise velocity

sensor to remove wave bias from Reynolds stress estimates by assuming equilibrium

44

Page 45: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

turbulence and no wave-turbulence interaction. Wave bias in a single-sensor Reynolds

stress estimate is calculated as:

~u ~w =fmax∑

fj =fmin

~U∗j

~Wj =fmax∑

fj =fmin

|~Uj || ~Wj | cos(

�Wj −�U

j

), (3–20)

where Uj and Wj are the Fourier transforms of u(t) and w (t), respectively, at frequency

fj . [fmin, fmax ] represents a frequency band dominated by waves, and is determined

by visual inspection of the spectra. ~|U j | and ~|W j | are obtained from the difference

between the spectra of the raw signals and the spectra obtained from the linear least

squares fit based on the assumption of the equilibrium turbulence. The phases of the

raw signal(

�Uj , �W

j

)are used to calculate the wave bias, because waves are expected

to dominate the spectrum of the raw signal under the wave peak. This simple method,

however, is likely to overestimate low-frequency turbulence in shallow water due to

infragravity wave bias, which is an effect of wave nonlinearities.

3.2.3 Two-sensor methods to reduce wave bias

Wave bias in Reynolds stress estimates can also be reduced by measuring

the flow field with two sensors, so that optimal distances can be found between the

measurement points and the bed, that will yield cross-correlated signals for waves and

uncorrelated signals for turbulence (Trowbridge, 1998). If the cross-sensor correlation

is low for turbulent fluctuations, but high for waves, cross-sensor differencing of the

measurements will not affect turbulent terms but will reduce the wave bias. Having

uncorrelated turbulent fluctuations at spatially separated points (i.e., two sensors) has

an empirical basis in the literature. The region over which turbulent fluctuations are

correlated is called an eddy, and this scale was found to be proportional to the height

of measurement above bed (Grant, 1958). Trowbridge and Elgar (2003) observed that

alongshore scales of turbulence contributing to near bottom Reynolds stresses range

from 0-4 times the height of measurement above the bottom. An empirical guideline was

presented by Shaw and Trowbridge (2001), based on the measurements obtained from

45

Page 46: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

a vertical array of five current-meters in the relatively wave-free lower Hudson estuary

(Trowbridge et al., 1999). The decay of turbulence correlation terms was calculated as a

function of vertical separation:

DT =Ruw (r )Ruw (0)

, (3–21)

where Ruw (r ) = u(zl ) w (zl + r ), r is the vertical separation of the two sensors,

and zl is the height of the lower sensor above the bottom. Shaw and Trowbridge

(2001) concluded that for the turbulence to be considered uncorrelated between

two sensors, i.e., DT less than 0.1, and to have the Reynolds stress estimates not

biased by turbulence correlation, r has to be greater than 5zl . Due to having only two

current meters in the vertical array and the presence of waves in the data, the decay

of turbulence correlation terms can not be checked for the data set analyzed here.

Therefore, the constraint given by Shaw and Trowbridge (2001) was followed in the

experimental setup (Section 2.2).

The two-sensor method requires an elaborate experimental setup, is most

effective for well-separated spatial scales of waves and turbulence, and is likely to

alias low-frequency turbulence as waves, but seems better suited for shallow ocean

environments. Based on this approach, the variance method (Lohrmann et al., 1990;

Stacey et al., 1999), which was initially proposed to estimate Reynolds stress profiles

using ADCPs in the absence of waves, has recently been modified to reduce wave-bias

in the estimates (Whipple et al., 2006; Rosman et al., 2008). The two-sensor method

(Trowbridge, 1998) is briefly explained here and the calculations are detailed in the

Appendix.

In the 2-D model coordinate system, let U1 and W1 be the horizontal and vertical

components of velocity, measured in instrument coordinates and accounting for their

small misalignment θ1 from the model coordinate system:

46

Page 47: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

U1 ≈ ~u1 + u′1 + θ1( ~w1 + w ′1) , (3–22)

W1 ≈ ~w1 + w ′1 − θ1(~u1 + u′1) . (3–23)

The subscript “1” (and subscript “2” given below) identifies the position of a measurement

or an estimate. The covariance of the measured velocities gives:

Cov(U1, W1) = u′1w ′1 + ~u1 ~w1 + θ1( ~w1

2 − ~u21) + θ1(w ′

12 − u′12) . (3–24)

The first term is the Reynolds stress at the sensor location, which is the quantity

of interest, the second and the third terms are wave biases, and the last term is a

turbulence bias. Assuming that all the components of the turbulent covariance tensor

are of the same order of magnitude (Tennekes and Lumley, 1972), and recalling that

θ1 is small, turbulence bias is smaller than the quantity of interest in equation (3–24).

However, wave biases need to be reduced in the presence of energetic wave motions.

By taking the difference of the velocity components measured at two spatially-separated

sensors, Trowbridge (1998) reduced wave contamination at the Reynolds stress

estimates, and calculated an average of Reynolds stresses at the two sensors:

u′1w ′1 + u′2w ′

2

2=

⟨u′w ′⟩ ≈ Cov {�U, �W }

2, (3–25)

where

�U = U1 − U2 ; �W = W1 −W2 . (3–26)

U2 and W2 are defined analogously to equations (3–22) and (3–23), but with subscripts

“1” replaced with “2”. Shaw and Trowbridge (2001) improved this raw differencing

method, such that wave bias can be reduced further by differencing after mapping the

horizontal component of velocity at one of the sensors, using the velocities measured at

47

Page 48: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

the other sensor, i.e., with linear filtration techniques. Also, the estimate at the sensor

location can be calculated, instead of an average of the Reynolds stresses at the two

sensor locations:

u′1w ′1 ≈ Cov

{�U, W1

}, (3–27)

where

�U = U1 − U1 . (3–28)

U1 is the horizontal velocity estimated at position “1” using the horizontal velocity

measured at position “2” with least squares linear adaptive filtering:

U1(t) =∞∫

−∞

h(t ′)U2(t − t ′)dt ′ , (3–29)

where h(t) is a filter that represents the relationship between the wave-induced

fluctuations at the two sensor locations. Here, low wave-bias estimates of Reynolds

stresses are obtained by the method presented by Feddersen and Williams (2007). This

method is actually a refinement of the approach developed by Trowbridge (1998) and

Shaw and Trowbridge (2001). Feddersen and Williams (2007) mapped both horizontal

and vertical components, and reduced the wave bias even further. The minus of the

product of density and the integral of the co-spectrum of the velocity differences gives a

nearly wave-free estimate of the Reynolds stress:

τ1 = −ρu′1w ′1 ≈ −ρCov(�U, �W ) = −ρ

∫ ∞

0Co�U,�W (f )df . (3–30)

Subscripts “1” and “2” specifically denote measurements of ADV-1 at 17 cmab and

ADV-2 at 145 cmab (Figure 2-5), respectively, hereinafter. For the vertical component of

velocity, W1 and �W are calculated through a procedure analogous to equations (3–28)

and (3–29). Co�U,�W (f ), the co-spectrum (equation (3–8)) of the velocity differences,

48

Page 49: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

is estimated based on the same Fourier analysis parameters as those used for wave

spectrum analysis (Section 3.1). This method can be used in principle to estimate

Reynolds stress at both ADV locations. However, away from the bed, the effect of the

bottom boundary decreases, and turbulent fluctuations become slower, increase in

scale, weaken, and are likely increasingly modulated by wave motion (Monismith and

Magnaudet, 1998; Bricker and Monismith, 2007). Therefore, Reynolds stresses are

estimated for the lower sensor (ADV-1 at 17 cmab, Figure 2-5).

The effectiveness of the two-sensor method (with differencing and filtering) in

reducing wave-bias varies with each measurement burst. Following atmosphere and

ocean boundary layer studies (Kaimal et al., 1972; Soulsby, 1977), Feddersen and

Williams (2007) proposed to accept as valid Reynolds stress estimates for which the

non-dimensional integrated co-spectrum (Og, also called ogive curve) satisfies the

condition:

− 0.5 ≤ Ogu′w ′(f ) =∫ f

0 Cou′w ′(s)dsu′w ′ ≤ 1.6 , (3–31)

for 0.1 < kN < 10, where kN = 2πfzV is the non-dimensional wavenumber, with z

the vertical distance above bed, and V the average along-stream velocity. Examples

of estimates considered relatively wave-biased or bias-free are shown in Figure 3-1.

Approximately half of the measurement bursts satisfy the condition in equation (3–31).

3.3 Logarithmic Law of the Wall

As mentioned in Section 1.3, floc formation is controlled by turbulent eddies which

are scaled by Kolmogorov micro-scale λ = 4√

ν3/ε (Berhane et al., 1997), where ν is the

kinematic viscosity of water, and ε is the dissipation rate of turbulent kinetic energy. ε is

related to the bottom friction velocity (u∗) as follows:

ε =u3∗

κ z, (3–32)

49

Page 50: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

where κ=0.41 is von Karman’s constant and z is a reference elevation above the bed.

For the investigation of the LISST data set from 2006, having u∗ estimates based on the

PC-ADP measurements from the same platform would allow the variation of floc size

distribution with a turbulent flow parameter to be seen. For this, the PC-ADP velocity

profiles are fit to logarithmic profiles, based on the logarithmic-law of the wall (called

log-law hereinafter). By following Nielsen (1992), the derivation of the log-law is shown

below, starting from the two-dimensional horizontal momentum equation:

∂u∂t

+ u∂u∂x

+ w∂u∂z

= −1ρ

∂p∂x

+ ν

(∂2u∂x 2 +

∂2u∂z 2

). (3–33)

The parameters in this equation are defined in this section previously. A steady flow (i.e.,

∂∂t =0) which is horizontal in the boundary layer (i.e., w ≈0) and horizontally uniform (i.e.,

u = u(z)) is assumed:

ν∂2u∂z 2 =

∂p∂x

. (3–34)

In addition to these assumptions, substituting the viscous shear stress τ = ρν ∂u∂z

simplifies equation (3–34) to:

∂τ

∂z=

∂p∂x

. (3–35)

Assuming hydrostatic pressure, i.e., the pressure gradient is constant over the depth:

∂τ

∂z=

∂p∂x

= ρg∂η

∂x. (3–36)

A bottom boundary condition is set by parameterizing the bottom shear stress as

τ(0) = τb = ρu2∗ . The velocity profile obeying the log-law in a turbulent boundary layer is

obtained as follows:

u(z)u∗

=1κ

ln(

zzo

), (3–37)

50

Page 51: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

where zo is a function of bottom roughness that describes a finite elevation above the

bed, where the velocity profile goes to zero.

The PC-ADP data set was collected in the presence of surface waves with

significant heights exceeding 1 m (Section 4.2). Since near-bed turbulence within

the wave boundary layer enhances bottom roughness experienced by currents (Grant

and Madsen, 1979, 1986), as discussed in Section 1.2, fitting a logarithmic profile to

the current velocity profiles in a least-squares sense (Lueck and Lu, 1997) would cause

errors in the resulting u∗ estimates. If there were not missing profiles in the PC-ADP

data set (Section 1.2), the method proposed by Rosman et al. (2008) could be applied

to estimate Reynolds stresses and then relate these to u∗. However, this method would

still require a modification due to different beam geometries of ADCPs and PC-ADPs.

Therefore, u∗ is estimated here from the mean current velocity profiles outside of the

wave boundary layer, following the approach of Lacy et al. (2005). The missing velocity

profiles also preclude obtaining accurate profiles of root-mean-square velocities, which

would give estimates about the thickness of the wave boundary layer. Therefore, a

constant thickness of 5 cm is assumed. The first three bins above the bottom are

assumed to be within the wave boundary layer (bin size was 1.6 cm, Section 2.2) and

not included in the analysis. With linear least-squares regression, a logarithmic profile

is fit to the velocity measurements at the first three bins just above this layer. A fit is

accepted to be valid if the correlation coefficient, r 2, between the measurements and the

fit is greater than 0.8. The bins above the first three bins are added into the logarithmic

fit, within 3-bin groups, as long as the resulting r 2 remains to be greater than 0.8.

51

Page 52: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

10−1

100

101

−0.5

0

0.5

1

kN

Og u’ w

Figure 3-1. Examples of ogive curves (cumulative integrals, equation (3–31)) as afunction of the dimensionless wavenumber kN . A low-bias estimate,corresponding to the 10-min measurement burst starting at 15:00 UTC onMarch 31st (thick line), and an estimate with significant wave-bias,corresponding to the burst starting at 05:00 UTC on April 1st (thin line).

52

Page 53: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

CHAPTER 4OBSERVATIONS

4.1 General Overview

At the experiment site, the water depth, averaged over the experiment duration, was

about 5 m (Figure 4-1a). Tidal ranges were mostly around 60 cm with an extreme of 80

cm, recorded close to the end of the experiment. Based on current-meter records

covering up to 30 months (Di Marco and Reid, 1998), and the results during the

2006 experiment in this area (Jaramillo, 2008), the dominant tidal constituents on

the Atchafalaya Shelf were identified to be K1, O1 and M2, with periods of 23.93 hrs,

25.82 hrs, and 12.42 hrs, respectively. Depth variation during this experiment was

controlled mainly by a diurnal signal, and the spectrum has a relatively weak peak near

the semidiurnal frequency, as well (Figure 4-1b). A more comprehensive tidal analysis is

beyond the scope of this study.

A 1-minute segment from the ADV measurements is shown in Figure 4-2. Good

synchronization of the instruments can be seen in all three of the velocity components,

especially in the horizontal velocity records (Figures 4-2a-c). The quality of the velocity

observations is illustrated in Figure 4-2d. Throughout the experiment, along-beam

correlations between successive acoustic returns from the scatterers in the water (Elgar

et al., 2005) stayed above 90% (Figure 4-2d), well within the acceptable limits, where the

lower limit recommended by the manufacturer is 70% (SonTek/YSI, 2001).

The first 4 days of the experiment (Figures 4-3 and 4-4, label 1) were characterized

by relatively calm weather. Winds exceeded 5 m/s only briefly at the beginning of

this period (Figure 4-3a), generating seas of about 0.5-m height (Figure 4-4a), and

decayed to about 3 m/s at the end of this period. Swell height never exceeded 0.5-m,

and remained below 0.3-m for most of this period.

The most energetic event observed during the experiment occurred on March

31st (Figures 4-3 and 4-4, label 2), characterized by steady winds blowing towards

53

Page 54: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

North-Northwest (Figure 4-3a). Wind speed reached 10 m/s for about a day on March

31st, forcing a detectable mean Northward-Northwestward current component (reaching

30 cm/s, Figure 4-3d). The atmospheric perturbation was likely local (relatively small

fetch) judging by the large seas generated (up to 1-m height, Figure 4-4a), but relatively

weak swell (about 0.25-m height).

The last part of the experiment (Figures 4-3 and 4-4, label 3) had the characteristics

of an atmospheric front passage, including the typical drop in wind velocity as the front

passed over the site (April 5th), and the rotation of wind and wave direction (onshore to

offshore) followed by the post-frontal wind intensification (Figures 4-3a and 4-3c). Wave

fields associated with this event also suggest a larger fetch: swells were the strongest

recorded during the experiment (Figure 4-3b), reaching up to 0.5-m height (Figure 4-4a)

and their arrival at the site was delayed with respect to the maximum winds.

The valid Reynolds stress estimates (Section 3.2.3) are evaluated with the

calculations based on the quadratic drag relation. The bottom drag coefficient is

calculated from its relation with the bottom stress, τb = ρu2∗ = ρCd

(u2 + v 2), equivalent

to Cd = u2∗/

(u2 + v 2). This can be rewritten separately for North-South and East-West

components as, τN−S = ρCdN−S u√

u2 + v 2, and τE−W = ρCdE−W v√

u2 + v 2 where u and

v are set hereinafter to denote North-South and East-West components, respectively.

Linear least square fits based on the Reynolds stress estimates (Figure 4-5) yield the

average drag coefficients for the entire duration of the experiment: CdN−S =2.3x10-3 with

a correlation coefficient of r 2=0.72 (Figure 4-5a), and CdE−W =2.9x10-3 with a correlation

coefficient of r 2=0.77 (Figure 4-5b), which are both within the standard expectations.

Figure 4-4e shows the response of near-bed flow and suspended sediment

concentration to surface flow forcing. The highest near-bed concentrations of suspended

sediment throughout the experiment were recorded during March 31st, with an average

of about 2 g/L and sometimes exceeding 3 g/L (Figure 4-4e, label 2). In general,

near-bed sediment resuspensions should respond to a number of forcing mechanisms,

54

Page 55: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

such as mean current velocity, current direction and divergence, and long-wave energy.

Short waves are typically not included in this list because of the strong attenuation

with depth. In the observations, seas consistently dominate swells in the surface

measurements throughout the experiment (Figure 4-4a). The two frequency bands

contribute almost equally to near-bed RMS orbital velocity (Figure 4-4b). However,

steeper short waves (seas) dominate near-bed accelerations (sometimes exceeding 0.5

m/s2) and exhibit a stronger correlation with sediment concentration values (compare

Figures 4-4c, 4-4d and 4-4e). This suggests that, for water depths characteristic for

this experiment, the short wave band is the dominant forcing mechanism that maintains

near-bed sediment suspensions. This is in agreement with previous observations and

numerical simulations of sand transport (Hoefel and Elgar, 2003; Hsu and Hanes, 2004).

Because the present observations were collected at one location in the horizontal

plane, it is not possible to assess the effect of horizontal advection on the observed

levels of suspended sediment concentration. The observed values of suspended

sediment concentration (Figure 4-4e) do follow the general trends in current velocity and

direction, increasing typically for stronger westward currents (Figure 4-3d). However,

due to the platform location (Figure 2-2), it is unlikely that the site was directly affected

by the sediment discharge at the mouth of the Atchafalaya River. Rather, several

characteristics of the observations seem to support the alternative hypothesis, that (on

the time and energy scales of this experiment) near-bed sediment suspensions are of

local origin and are maintained by local wave-induced turbulence. Bottom core samples

collected before and after the experiment suggest an effectively inexhaustible local

reserve of mobilizable, soft bottom sediment. Suspended sediment concentration values

are well correlated with the wave-induced accelerations (Figure 4-4c). The evolution

of Reynolds stress estimates and suspended sediment concentration (Figure 4-6) is

consistent with the assumption that local turbulence is the main forcing mechanism

driving near-bed suspensions.

55

Page 56: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

The strong correlation between Reynolds stresses and suspended sediment

concentration in Figure 4-6 suggests a nearly-monotonic relation between the two

parameters and, implicitly, a weak dependence of the turbulence-sediment processes

on the third parameter – the floc size. This unexpected result could be an expression of

a general weak dependence on the floc size; however, the available data set does not

cover a wide enough range of concentrations to support this. Instead, we adopt here

the conservative interpretation that, for the relatively dilute suspension conditions in

this experiment, floc size variability is small enough to justify neglecting its effects on

turbulence-cohesive sediment interactions.

4.2 The LISST Data Set from 2006

This hypothesis about floc size variation is supported by the available information

about suspended sediment size distribution on the Atchafalaya Shelf, i.e., LISST

observations, and the co-located PC-ADP measurements of waves and currents,

collected in 2006 (Figure 2-2). A comprehensive analysis of the near-bed flow

measurements of the PC-ADP was done by Jaramillo et al. (2009). For the analysis

herein, only the mean current profiles are analyzed to estimate the bottom friction

velocity, and the significant wave heights are calculated from pressure spectrum (with

17 degrees of freedom). Figure 4-7 shows the significant heights of seas and swells

(Figure 4-7a), vertical structure of current speeds (Figure 4-7b), and the estimates

of bottom friction velocity (Figure 4-7c), based on log-law (Section 3.3). For 82% of

the measurement bursts, valid logarithmic layers (r 2 greater than 0.8, Section 3.3)

are obtained. While the vertical range of the estimated logarithmic layers vary (black

“x”s in Figure 4-7b), a logarithmic layer is fit to almost the entire profiling range of

the PC-ADP for majority of the cases. The average r 2 of the valid logarithmic fits is

0.97, with more than 90% of them having r 2 greater than 0.9. Therefore, for this data

set, the log-law assumption describes the near-bed current structure very well. Two

examples of logarithmic fits to the current profiles are shown in Figure 4-8a, together

56

Page 57: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

with the corresponding LISST records. Size distribution is given as fractional volume of

particles in µL/L. The distributions for both of these cases have a major peak between

200-230 µm and a smaller peak near 50µm, which is in agreement with the general

trend throughout the experiment, as discussed below.

Throughout the experiment, the size distribution exhibits two modes (Figure 4-9a).

The dominant mode is within the 150-300 µm range, with the peak position between

200-230 µm. A weaker mode, which is centered around 50 µm, appears to be sensitive

to the amount of suspended sediment in the water column. The position of the peaks

of two modes are remarkably stable, given the varying bottom turbulence (Figure 4-9b)

and surface wave conditions (Figure 4-9c). The dominant grain-size is in the range of

coarse sand. However, because there is not coarse sand in this region, it is likely to

represent the floc mode. The weaker mode is probably due to either coarse silt particles

in suspension or smaller flocs.

To examine the bottom turbulence and cohesive sediment interaction in more

detail, and the effects and uncertainties due to floc size (i.e., settling velocity), a bottom

boundary layer model for cohesive beds is used in the next section. The fundamental

assumption above is critical in allowing for numerical analysis, as existing numerical

models of flow-sediment interaction cannot yet handle floc aggregation-breakup

processes. In this study, the floc size will be assumed to vary on a time scale longer

than the 10-min duration of a measurement burst. Therefore, it will be approximated as

constant for the duration of a measurement burst, but it will be assumed to vary slowly

from burst to burst.

57

Page 58: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

03/28 04/02 04/074.4

4.6

4.8

5

5.2

5.4D

epth

(m

)

time (month/day 2008)

(a)

0 1 2 3 40

0.01

0.02

0.03

Spe

ctra

lde

nsity

(m

2 /cpd

)

Frequency (cpd)

(b)

Figure 4-1. Tidal variations during the experiment. (a) Time evolution of the mean waterdepth at the measurement location; (b) Fourier spectrum of the depthvariations, with 10 degrees of freedom.

58

Page 59: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

0

0.2

0.4N

orth

−co

mpo

nent

of v

eloc

ity (

m/s

)(a)

−0.2

−0.1

0

0.1

Eas

t−co

mpo

nent

of v

eloc

ity (

m/s

)

(b)

−0.1

0

0.1

Ver

tical

−co

mpo

nent

of v

eloc

ity (

m/s

)

(c)

5 10 15 20 25 30 35 40 45 50 55 6090

95

100

Cor

rela

tion

(%)

(d)

time (second)

Figure 4-2. A one-minute segment of the flow velocity, recorded by the ADV array. (a)North component of velocity; (b) East component of velocity; (c) verticalvelocity; and (d) 3-beam average of the ADV signal correlation. Thin andthick lines show the records of ADV-1 at 17 cmab and ADV-2 at 145 cmab,respectively. The measurements correspond to the first minute of the burst,starting at 11:00 UTC on March 30th, 2008.

59

Page 60: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

Figure 4-3. Wind, wave, and current conditions on the Atchafalaya Shelf throughout the2-week experiment. Time evolution of: (a) wind speed and direction(color-coded); (b) spectral density, normalized between 0 and 1, based onthe ADCP measurements; (c) peak wave propagation direction for eachfrequency band in the power spectrum, based on the ADCP measurements(directions are shown only for frequencies with spectral density greater than10-2 m2/(Hz.deg)); and (d) current speed and direction at 17 cmab (ADV-1).The direction indicated is the direction of the flow (e.g., N means flowingnorthward).

60

Page 61: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

0

0.5

1

Wav

ehe

ight

(m

) (a)

0

0.05

0.1

0.15

RM

S

velo

city

(m

/s) (b)

0

0.2

0.4

RM

Sac

cele

ratio

n (m

/s2 )

(c)

0

0.02

0.04

0.06

Ste

epne

ss (d)

03/27 03/29 03/31 04/02 04/04 04/060

1

2

3

Sus

pend

ed s

edim

ent

conc

entr

atio

n (g

/L)

time (month/day 2008)

(e)

1 2 3

Figure 4-4. Bulk spectral characteristics of waves, and the measurements of suspendedsediment concentration. Time evolution of: (a) significant wave height at thesurface; (b) RMS velocity estimated at 17 cmab (ADV-1); (c) RMSacceleration estimated at 17 cmab (ADV-1); (d) surface wave steepness; and(e) suspended sediment concentration (10-min averages) measurements at12 cmab (OBS-5). In panels a,b,c and d, swell and sea frequency bands aredenoted by a thick line and a thin line, respectively.

61

Page 62: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

−50 0 50 100

−0.2

−0.1

0

0.1

0.2

0.3τ

nort

h (P

a)

(a)

ρu

u2

+ v2

−100 −50 0−0.5

−0.4

−0.3

−0.2

−0.1

0

0.1

τ ea

st (

Pa)

(b)

ρv

u2

+ v2

Figure 4-5. Reynolds stress estimates versus the quadratic drag relation. (a)North-South components (CdN−S =2.3x10-3, r 2=0.72); (b) East-Westcomponents (CdE−W =2.9x10-3, r 2=0.77). The straight lines show the linearleast square fits.

62

Page 63: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

0

0.1

0.2

0.3

0.4

Rey

nold

sst

ress

(N

/m2 ) (a)

03/27 03/29 03/31 04/02 04/04 04/060

0.5

1

1.5

2

2.5

3

3.5

(b)

Sus

pend

ed s

edim

ent

conc

entr

atio

n (g

/L)

time (month/day 2008)

1 2 3

Figure 4-6. Reynolds stress estimates, and the measurements of suspended sedimentconcentration. Time evolution of: (a) the magnitudes of the Reynolds stressestimates (filled circles and squares denote magnitudes of North-South andEast-West components, respectively); and (b) suspended sedimentconcentration.

63

Page 64: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

03/02 03/04 03/06 03/08 03/10 03/12 03/140

0.01

0.02

0.03

0.04

0.05

(c)

Bot

tom

fric

tion

velo

city

(m

/s)

time (month/day 2006)

0

0.5

1(a)

Wav

e he

ight

(m

)D

ista

nce

from

se

nsor

hea

d (m

)

(b)

0.2

0.4

0.6

0.8

1

Cur

rent

spee

d (m

/s)

0

0.2

0.4

Figure 4-7. Wave and current observations during the 2006 experiment. (a) Significantwave height at the surface in the sea (f >0.2 Hz, thin line) and swell (f ≤0.2Hz, thick line) bands; (b) vertical profiles of mean horizontal current speeds(“x”s indicate the elevations where the logarithmic-layer of the profilesreach); and (c) bottom friction velocity estimates based on the log-law.

64

Page 65: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

0 0.1 0.2 0.3

0

0.2

0.4

0.6

0.8

Dis

tanc

e ab

ove

botto

m (

m)

(a)

0 0.1 0.2 0.30

0.2

0.4

0.6

0.8

Current speed (m/s)

Dis

tanc

e ab

ove

botto

m (

m)

(c)

0 50 100 150 200 250 300

0

5

10

15

Fra

ctio

nal

volu

me

(µL/

L)

(b)

0 50 100 150 200 250 3000

5

10

15

Particle size (µm)

Fra

ctio

nal

volu

me

(µL/

L)

(d)

Figure 4-8. Examples of logarithmic layer fits, and grain size distributions from the 2006experiment. In panels a and c, circles denote the PC-ADP measurements ofcurrent speeds, and the thick lines show the logarithmic fits to themeasurements outside of the wave boundary layer. In panels b and d, sizedistribution records of LISST are given in fractional volume of each particlesize. Panels a and b correspond to the burst starting at 06:00 UTC on March4th, panels c and d correspond to the burst starting at 12:00 UTC on March8th.

65

Page 66: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

Par

ticle

si

ze (

µm)

(a)50

100

150

200

250

300

log 10

(Fra

c.vo

l.)

(µL

/L)

0

1

2

03/02 03/04 03/06 03/080

0.5

1

W

ave

heig

ht (

m)

(c)

time (month/day 2006)

0

0.01

0.02

0.03

Bot

tom

fric

tion

velo

city

(m

/s)

(b)

Figure 4-9. Observations of particle size distribution and wave-turbulence conditions in2006. (a) Fractional volume of each particle size, calculated fromhourly-averaged LISST data; (b) estimates of bottom friction velocity; and (c)significant wave height at the surface in swell (thick line) and sea bands (thinline).

66

Page 67: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

CHAPTER 5BOTTOM BOUNDARY LAYER MODELING

5.1 Introduction

Analysis of turbulent flow interacting with cohesive sediments can be strengthened

by numerical modeling, since flow-related parameters, which can not be directly

measured or estimated, can be calculated. The numerical analysis of the observations

herein is based on the one-dimensional bottom boundary layer model developed by Hsu

et al. (2007) for cohesive beds and improved by Hsu et al. (2009). The model assumes

a constant floc size (also constant floc density and settling velocity) and integrates the

two-phase (fluid-sediment) Reynolds-averaged equations based on a turbulent kinetic

energy (TKE, k , equation (3–14)) - dissipation rate of turbulence kinetic energy (ε)

closure. The model is different than a single-phase model in the sense that sediment is

not considered to be passive, but the effect of sediment on fluid turbulence is accounted

for (Hsu and Liu, 2004). The details of the numerical model and the derivations can

be found in Hsu et al. (2007). The governing equations for momentum, sediment

concentration, k − ε balances, and the boundary conditions are cited in Section 5.2

for convenience. Then, numerical simulations are done for the measurement bursts,

which were collected on March 31st 2008 (within event 2 in Figures 4-3, 4-4, and 4-6)

and provided valid Reynolds stress estimates (Sections 3.2.3 and 4.1, Figure 4-6a).

This is the interval when the most energetic flow conditions were observed and these

conditions caused the amount of suspended sediment to reach its maximum values

throughout the experiment.

5.2 Governing Equations

In order to make the presentation of the governing equations simpler, the numerical

coefficients in the k − ε closure are given in advance (Table 5-1), and the overbars that

represent ensemble-averaged quantities in the equations are dropped.

67

Page 68: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

5.2.1 Momentum balance

The coordinate system is defined such that x and y are the horizontal directions

and z is the vertical direction normal to the bed. The cohesive fine sediments of

interest were noted to have particle response time of O(0.001 sec), which is very

small compared to the timescale of flow forcing, therefore the sediment is assumed to

follow the fluid velocity (Hsu et al., 2007). Therefore, the two-phase flow velocity can be

described by a single mean flow velocity, i.e., one momentum equation is necessary to

describe the flow. The horizontal momentum equations (equations 1 and 2 in Hsu et al.

(2007)) read:

∂uj

∂t=

−1ρ(1− φ)

∂p∂xj

+1

ρ(1− φ)

[∂τw

jz

∂z+

∂τ sjz

∂z

]+ gj

[(s − 1)φ

1− φ

], (5–1)

where uj is the j-th (j=1,2) horizontal component of velocity along the horizontal

dimension xj , φ is the volume concentration of sediment floc, τjz is the element of the

stress tensor, and gj is the j-th horizontal component of the gravitational acceleration.

The superscripts “s” and “w” denote the quantities corresponding to sediment and

water, respectively, and s = ρm/ρ is the specific gravity of cohesive sediment floc with

density ρm. The right-hand side terms in equation (5–1) describe forcing by free-stream

pressure gradient in the horizontal, momentum transport caused by fluid and sediment

shear stresses (including both laminar and turbulent components), i.e., diffusion, and

transport of momentum by gravity-driven fluid flow. The model assumes a horizontally

uniform bottom boundary layer, therefore the advective acceleration terms are neglected

in this formulation. Due to the mild slope at the experiment site which is about 10-3, the

implementation of the equation (5–1) for this study neglects the terms of gravity-driven

transport of momentum. Horizontal pressure gradient is defined as the prescribed flow

forcing of the bottom boundary layer, i.e., velocity time series which includes oscillatory

forcing and a mean current component as follows (equation 23 in Hsu et al. (2007)):

68

Page 69: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

− ∂p∂xj

= ρ∂ �Uj (t)

∂t+ f c

j , (5–2)

where �Uj (t) is the de-meaned time-series of velocity which represents oscillatory forcing

and prescribed into the model directly as input. As addressed in Section 5.4, an iterative

procedure is followed to calculate the forcing f cj (function of the bottom friction velocity

u∗j ), that is required to simulate the measured currents. The fluid shear stress in the

equation (5–1) is calculated as:

τwjz = ρ (ν + νt)

∂uj

∂z, (5–3)

where νt is the eddy viscosity, given as:

νt = Cµk2

ε(1− φ) . (5–4)

Calculations of k and ε are given in Section 5.2.3. The sediment shear stress in equation

(5–1) is calculated as:

τ sjz = ρνr

∂uj

∂z. (5–5)

νr is the relative viscosity, and is calculated as a function of the sediment volume

concentration (Zarraga et al., 2000):

νr = νexp(−2.34φ)(1− φ/φo)3 , (5–6)

where φo is the gelling concentration for mud flocs (Winterwerp and van Kesteren, 2004)

at which a space filling network develops and settling velocity becomes zero.

The vertical velocity shear (e.g., equations (5–3) and (5–5)) is calculated by

following log-law. As the bottom boundary condition, the shear stress at the bed (z = 0)

is calculated as τjz (0) = ρu2∗j .

69

Page 70: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

5.2.2 Sediment concentration balance

The evolution of sediment concentration is given as (equation 3 in Hsu et al.

(2007)):

∂φ

∂t= − ∂

∂z

[φ(1− φ)Tp(1− s−1)g − νt

σc

∂φ

∂z+

Tp

ρs

∂τ szz

∂z

], (5–7)

where Tp is the particle response time. The first term on the right-hand side describes

gravitational settling. The second term is the closed form of turbulent mass flux,

which is, in fact, the covariance of concentration fluctuations and vertical velocity

fluctuations, i.e., φ′w ′. The last term in equation (5–7) is the vertical gradient of sediment

intergranular normal stresses, i.e., rheology, which serves as an additional suspension

mechanism. For the low values of suspended sediment concentration measured

(Figures 4-4e and 4-6b), the resuspension effect of rheological stress is assumed to be

negligible. Therefore, suspended sediment concentration is established by a balance of

gravitational settling and turbulent mass flux.

The bottom boundary condition for downward flux is approximated by continuous

deposition. Upward flux at the bottom boundary is determined by defining a resuspension

coefficient (γo), and a critical shear stress to initiate sediment motion (τc ). The

resuspension coefficient controls sediment availability from the bed and has to be

defined, together with the critical shear stress, in the numerical simulation process by

matching measured and simulated suspended sediment concentration values. More

advanced erodibility formulations (e.g., Stevens et al., 2007) are not used in this study,

as they require information about bed erodibility parameters that are not available for this

experiment.

5.2.3 Turbulent kinetic energy balance and balance of turbulent kinetic energydissipation rate

The equations for turbulent kinetic energy (k) and its dissipation rate (ε) are

(equations 9 and 10 in Hsu et al. (2007)):

70

Page 71: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

(1− φ)∂k∂t

= νt

[(∂u∂z

)2

+(

∂v∂z

)2]

+∂

∂z

[(ν +

νt

σk

)∂(1− φ)k

∂z

]− (1− φ)ε

− (s − 1) gνt

σc

∂φ

∂z− 2φsk

Tp + TL, (5–8)

and

(1− φ)∂ε

∂t= Cε1

ε

kνt

[(∂u∂z

)2

+(

∂v∂z

)2]

+∂

∂z

[(ν +

νt

σε

)∂(1− φ)ε

∂z

]

−Cε2ε2

k(1− φ) + Cε3

ε

k

[− (s − 1) g

νt

σc

∂φ

∂z− 2φsk

Tp + TL

], (5–9)

where TL is the turbulent eddy timescale. The first three right-hand side terms in

equations (5–8) and (5–9) describe shear-induced turbulence production, diffusion,

and turbulent dissipation. The fourth terms represent the damping effect of suspended

sediment on fluid turbulence by density-induced stratification. This term accounts for

the vertical gradient of suspended sediment concentration, i.e., density throughout

the water column, which reduces vertical turbulent mixing due to buoyancy flux within

the stratified layer. The stratification effects are discussed in detail in Sections 5.4 and

5.5. The last terms in equations (5–8) and (5–9) represent the effect of suspended

sediment on turbulence by fluid-sediment velocity interaction through viscous drag.

For the low values of suspended sediment concentration measured (Figure 4-6b), the

sediment effects on turbulence due to viscous drag are assumed negligible. This leaves

stratification as the only damping effect of sediment in turbulence closure.

At the bottom boundary, a no-flux condition is set for k , and a standard near-wall

approximation is made for ε (Hsu et al., 2007):

∂k∂z

= 0 ; ε =C

34µ k

32

κz. (5–10)

71

Page 72: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

5.2.4 Sediment definition

The sediment phase is defined by a primary particle size (Dp) with density

ρs = 2.65ρ, a floc size (D), and the fractal dimension of the floc (nf ), all of which are

independent of time and position. Density (ρm) (Kranenburg, 1994), mass concentration

(c), and settling velocity (ws) of flocs are then calculated as:

ρm = ρ +(

Dp

D

)3−nf

(ρs − ρ) , (5–11)

c = ρsφ

(Dp

D

)3−nf

, (5–12)

ws =(s − 1)D2g

18ν(1− φ)q . (5–13)

In the presentation of the model, q was given equal to 4. The selection of floc size

input is based on the LISST data (Section 4.2), and the results of a flocculation model

(Section 5.3) that uses the Reynolds stress estimates (Section 3.2.3, Figure 4-6a) and

the sediment concentration measurements (Section 4.1, Figure 4-6b) as inputs.

5.3 Analytical Flocculation Model

Winterwerp (1998) proposed a model to calculate an equilibrium floc size and

settling velocity for given turbulence conditions and sediment information, for a constant

fractal dimension of 2. In the model, flocculation is described by a linear combination of

formulations for simultaneous aggregation and floc breakup due to turbulent motion. The

aggregation is formulated by assuming that eddies larger than the Kolmogorov scale

bring particles smaller than the Kolmogorov scale together (Levich, 1962). The fact that

not all collisions will result in flocculation is included in the model through an empirical

coefficient in the aggregation formulation. The effects of collision between particles due

to Brownian motion and overtaking of particles with small settling velocity by those with

large settling velocity, i.e., differential settling, are neglected. Floc breakup by turbulent

72

Page 73: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

shear is formulated as a function of the ratio of the shear stress and the floc strength. An

empirical coefficient is implemented in the formulation due to the uncertainties related to

the floc strength. The formulation for simultaneous aggregation and floc breakup is given

as:

dDdt

= kAcaveGD2 − kBG 3/2D2 (D −Dp) , (5–14)

where kA and kB are calibration coefficients, cave is the depth-averaged suspended

sediment concentration, and G is the dissipation parameter, which is a measure of the

turbulent shear in the flow. G is related to the shear stress, such that τ = ρνG in the

viscous regime at the Kolmogorov scale. The terms on the right hand side of equation

(5–14) are growth and breakup terms, which scale with D2 and D3, respectively.

Therefore, the growth term dominates for small flocs, and the breakup term dominates

for large flocs. The equilibrium floc size (De) is obtained for dDdt = 0 as:

De = Dp +kAcave

kB√

G. (5–15)

For details of the derivation, see Winterwerp (1998). Winterwerp (1998) obtained the

calibration coefficients from a series of laboratory experiments. In a settling column,

sediment samples with Dp=4 µm and varying concentration were mixed homogeneously.

Then, a homogeneous turbulence field was generated in the settling column by a grid,

oscillating at varying frequencies, i.e., varying G . Based on the floc sizes estimated

for the sediment samples withdrawn from the column, Winterwerp (1998) obtained the

empirical coefficients in equations (5–14) and (5–15) as kA=14.6 m2/kg and kB=14x103

s1/2/m2.

Through equation (5–15), equilibrium floc sizes are predicted for the simulated

bursts (Section 5.4). The primary particle size, representative for the experiment site,

is taken as Dp=5 µm (Section 2.1). The measured concentrations, and the estimated

Reynolds stresses for these bursts are used as inputs. Equation (5–15) yields floc size

73

Page 74: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

values between 73 µm (ρm=1.14 g/cm3) and 335 µm (ρm=1.05 g/cm3), with an average

of 182 µm (ρm=1.07 g/cm3). These floc densities are calculated with nf =2, following the

assumption made by Winterwerp (1998). The calculated floc sizes are consistent with

the dominant peak in the LISST observations of grain-size distribution (Section 4.2,

Figure 4-9a). The reader is cautioned, however, that equation (5–15) is an idealized

model, and has limited applicability in a field study. Therefore, consistent with the LISST

observations (Section 4.2, Figure 4-9a), a representative constant floc size D=200 µm

is used for all the numerical simulations, with a fractal dimension of nf =2.3 (Khelifa and

Hill, 2006). This yields ρm =1.15 g/cm3 (equation (5–11)) for a primary particle diameter

Dp=5 µm (Section 2.1). The resulting sediment settling velocity, based on Stokes’ law

(equation (5–13)), is ws=3.3 mm/s.

5.4 Application

The model domain is set to extend between the bed and the sampling volume

of ADV-2 at 145 cmab. At the top of the domain, the model is constrained by the

de-meaned horizontal velocity measurements. A trial and error process is used to

estimate the components of the friction velocity and resuspension coefficient (γo) for

which the model predictions of the mean flow at 145 cmab and suspended sediment

concentration at 12 cmab (OBS-5 location) match the observations. Previous numerical

studies, based on the model used here (Hsu et al., 2009), show that wave-averaged

values of suspended sediment concentration are not sensitive to the critical shear stress

(τc ), because the variation in the critical shear stress is compensated by the variability of

the resuspension coefficient during the model initialization procedure. The simulations

presented here used a constant critical bottom shear stress of 0.4 Pa, within the range

of 0.05-1.0 Pa investigated by Hsu et al. (2007, 2009). In the numerical experiments

with several τc -γo pairs, the model results did not indicate a sensitivity on τc in this

formulation. The procedure gradually “spins up” the model from a zero-flow state to

a steady state (such that the Reynolds-averaged flow parameters become constant

74

Page 75: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

in several consecutive runs with the same set of inputs), achieved typically within

a time duration that scales with the model domain height and eddy viscosity, which

controls the mixing time. For a model domain of 1 m, and νt= 10−4-10−3 m2/s, the time

necessary to reach a steady state is of the order of 103-104 sec in model time. Below,

we discuss numerical simulations for 13 of the most energetic 10-min measurement

bursts, collected on March 31st (within event 2 in Figures 4-3, 4-4, and 4-6) with low

wave-bias Reynolds stress estimates.

Examples of time series of model simulations are shown in Figure 5-1, for the

10-min measurement burst starting at 11:00 UTC on March 31st. Figures 5-1a and b

show the North-component of velocity simulated at the top of the model domain (dashed

line in Figure 5-1a denotes the northward current at 145 cmab, the location of the

sampling volume of ADV-2), and the suspended sediment concentration simulated at the

lower 20 cm of the model domain, respectively. The numerical results match observed

values: suspended sediment concentrations at 12 cmab, the location of the OBS-5

(Figure 5-1c) agree well with “sub-burst” (1-min average) OBS-5 observations (thick line

in Figure 5-1c). This agreement is remarkable given that the model is “spun up” only

through matching the 10-min averaged values, and shows that the model captures the

small-scale trends of the process. The North-component of the Reynolds stress at 17

cmab, the location of the sampling volume of ADV-1 (Figure 5-1d), is of the same order

as the estimate based on the observations (dashed line in Figure 5-1d). A more detailed

analysis of the performance of the model can be found in Hsu et al. (2007, 2009).

Figure 5-2 compares the calculations of the numerical model with measurements

and Reynolds stress estimates. The model reproduces the 10-min averages of

suspended sediment concentration observed at the OBS-5 location (Figure 5-2a)

and the mean currents measured at the locations of the two ADVs (Figures 5-2b and

5-2c). Suspended sediment concentration, vertically-averaged over the model domain,

varies between 0.44-1.41 g/L, well above the threshold 0.1 g/L suggested by Winterwerp

75

Page 76: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

(2001) and Winterwerp (2006) for sediment-induced stratification effects to modify

the turbulence field. The agreement between the model and the estimated Reynolds

stresses (Figures 5-2d and 5-2e), albeit less good, is encouraging, considering the

involved processing and the multi-layered hypotheses on which both the model

results and the estimates are separately based. The model agrees better with the

estimates of the West components of the Reynolds stress (Figure 5-2e) probably due to

a higher mean to fluctuation ratio for the West velocity component (Figure 5-2f), usually

translating into a lower wave-bias estimate (Feddersen and Williams, 2007).

The overall model-observations agreement justifies a more detailed analysis of

the model representation of the bottom turbulence-cohesive sediment interaction.

The vertical structures of the four terms in the governing TKE equation (5–8) are

calculated. Figure 5-3a shows Reynolds-averaged vertical profiles of four terms in the

governing TKE equation (5–8) calculated by the model for a 10-min measurement burst

characterized by one of the highest suspended sediment concentrations (2.83 g/L)

observed during the experiment. In the upper 10 cm of the model domain, vertical shear

due to velocity is relatively small and therefore TKE production is mainly due to turbulent

diffusion (stars in Figure 5-3a) balanced by sediment-induced stratification (“x”s in

Figure 5-3a). Below this upper layer of the model domain, the shear-induced turbulence

production (circles in Figure 5-3a) becomes the dominant turbulence generating term

due to increasing bottom boundary effect. At 100 cmab, sediment-induced stratification

(“x”s in Figures 5-3a and 5-3b) is the dominant turbulence damping term in most of

the cases, however, shear-induced turbulence production is balanced by the combined

effects of sediment-induced stratification and turbulent dissipation (Figures 5-3a and

5-3b). At 30 cmab, these two effects are of the same order of magnitude (“x”s and

squares in Figures 5-3a and 5-3c). Below 10 cmab, just above the bed, the TKE

balance is dominated by shear-induced turbulence production and turbulent dissipation.

Although overall sediment-induced stratification damping decreases with increasing

76

Page 77: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

distance above the bed, because the other terms are decreasing faster, paradoxically,

the role of stratification in the TKE balance becomes more important with increasing

distance above the bed. This suggests that, for the conditions of this experiment,

sediment-induced stratification term can be neglected in the wave boundary layer, but is

essential within the tidal boundary layer. Compared to the other three terms in the TKE

balance, vertical variation of sediment-induced stratification is smaller, as it varies within

two orders of magnitude throughout the model domain (Figure 5-3a).

One quantity of interest that the numerical model calculates is the turbulent mass

flux term φ′w ′ (equation A11 in Hsu et al. (2007)) in the sediment concentration

balance (5–7). In principle, this quantity could be estimated from the observations

by calibrating the acoustic backscatter records of an ADV with co-located suspended

sediment concentration measurements (e.g., Fugate and Friedrichs, 2002). However,

uncertainties due to application of this method in the presence of waves, and not having

the ADV-1 and the OBS-5 sampling at the same vertical level, make it impossible to

control the errors of such an estimate.

5.5 Model Sensitivity to Floc Size

To investigate the sensitivity of the model to floc size, additional numerical

simulations were carried out for one burst, with floc sizes of D=150 µm (ρm=1.18

g/cm3, ws =2.2 mm/s), and D=300 µm (ρm=1.12 g/cm3, ws=5.9 mm/s), while keeping

all other parameters constant (the resuspension coefficient γo varies in each test

run to match the model and the measured suspended sediment concentration). In

the LISST measurements (Figure 4-9a), the interval [150,300] µm brackets the main

distribution mode and has a time-averaged probability of 85%. The results (Figure 5-4)

suggest that decreasing floc size (i.e., decreasing settling velocity) decreases near-bed

suspended sediment concentration, produces a well-mixed concentration profile (Figure

5-4a), and increases the total amount of suspended sediment. The corresponding

vertically-averaged suspended sediment concentration values are 1.37 g/L for 150 µm,

77

Page 78: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

1.15 g/L for 200 µm, and 0.96 g/L for 300 µm. This is consistent, for example, with the

saturation concentration proposed by Winterwerp (2001) (equation 3 therein). Mean

currents are largely insensitive to floc size variation between 150 µm and 200 µm flocs,

however, 300 µm flocs result in a slight over-prediction of currents at 17 cmab (Figures

5-4b and c). The vertical structure of Reynolds stresses is not modified significantly

and, therefore, not shown here due to overlapping vertical profiles. To see the effect

of floc size variation on stratification effects, the gradient Richardson number (Ri )

associated with sediment-induced stratification is calculated. This is the ratio between

the stratification term and shear-induced turbulence production term in equation (5–8):

Ri = −(s−1)g

σc

∂φ∂z[(

∂u∂z

)2 +(

∂v∂z

)2] . (5–16)

The vertical structure of the gradient Richardson number is similar for all three

floc sizes tested (Figure 5-4d). It exceeds the critical value of 0.25 above a height

of 5-20 cmab (depending on the floc size), and then increases slowly with height

above the bed, slightly faster for the smaller floc size (i.e., for higher total amount of

suspended sediment). Relatively stratified profile near bed, obtained with larger flocs,

causes stratification to be important, i.e., Ri >0.25, in a wider domain (dashed lines in

Figures 5-4a and d). This increase in turbulence damping enhances mean current flow

throughout the model domain (dashed lines in Figures 5-4b and c) which is known as

drag reduction. The higher values of Ri at the top of the model domain are due to the

boundary condition of zero velocity shear at the top of the domain.

78

Page 79: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

Table 5-1. Numerical coefficients in the k − ε closureCµ Cε1 Cε2 Cε3 σk σε σc

0.09 1.44 1.92 1.20 1.00 1.30 1.00

79

Page 80: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

0

0.2

0.4

Nor

th−

com

pone

ntof

vel

ocity

(m

/s)

(a)

Dis

tanc

eab

ove

bed

(cm

)

(b)5

10

15

20

log 10

(SS

C)

(g/L

)

0.4

0.8

1.2

2

2.5

3

3.5

4

Sus

pend

ed s

edim

ent

conc

entr

atio

n (g

/L)

(c)

0 1 2 3 4 5 6 7 8 9 100.1

0.15

0.2

0.25

0.3

Nor

th−

com

pone

ntof

Rey

nold

s st

ress

(P

a)

time (minute)

(d)

Figure 5-1. An example of time-series in the model simulations. (a) North-component ofvelocity at 145 cmab (ADV-2 location); (b) suspended sedimentconcentration at the lower 20 cm above bed; (c) suspended sedimentconcentration at 12 cmab (OBS-5 location); and (d) the North-component ofReynolds stress at 17 cmab (ADV-1 location) for the 10-min measurementburst which started at 11:00 UTC on March 31st. In panel (a), the dashedline represents the North-component of the (10-min averaged) currentmeasured at 145 cmab. In panel (c), thick lines denote the 1-min time spansof the OBS-5 measurements of suspended sediment concentration. In panel(d), the dashed line marks the value of the (10-min averaged) estimate forthe North component of Reynolds stress at ADV-1.

80

Page 81: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

1 2 30.5

1

1.5

2

2.5

3

3.5

Sus

pend

ed s

edim

ent

conc

entr

atio

n−m

odel

(g/

L)

Suspended sediment concentration−meas (g/L)

(a)

0.1 0.2 0.3 0.4

0.1

0.2

0.3

0.4

|u| m

odel

(m

/s)

|u|meas

(m/s)

(b)

0 0.2 0.40

0.1

0.2

0.3

0.4

0.5

|v| m

odel

(m

/s)

|v|meas

(m/s)

(c)

0 0.2 0.40

0.1

0.2

0.3

0.4

|τu−

mod

el| (

Pa)

|τu−est

| (Pa)

(d)

0 0.2 0.40

0.1

0.2

0.3

0.4

0.5|τ

v−m

odel

| (P

a)

|τv−est

| (Pa)

(e)

0.06 0.08 0.10.05

0.1

0.15

0.2

0.25

0.3

σx (m/s)

|x| (

m/s

)

(f)

Figure 5-2. Comparison of observations and model results. (a) Suspended sedimentconcentration (at 12 cmab, the location of OBS-5, Figure 2-5); (b) Northcomponent of current velocity (“x” – 17 cmab, location of ADV-1; circles –145 cmab, location of ADV-2, Figure 2-5); (c) West component of currentvelocity (same symbols as in b); (d) North component of Reynolds stress at17 cmab; (e) West component of Reynolds stress at 17 cmab; (f) mean vs.standard deviation of velocity measurements at 17 cmab (crosses – Northcomponent; stars – West component).

81

Page 82: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

10−7

10−6

10−5

10−4

10−3

0

20

40

60

80

100

120

140(a)

D

ista

nce

abov

e be

d (c

m)

Contribution to TKE balance (m2/s3)

0 3 6 9 12 15 18 21

10−6

10−5

10−4

Con

trib

utio

n to

TK

E b

alan

ce (

m2 /s

3 )

(c)

time (hour on March 31st)

10−7

10−6

10−5

Con

trib

utio

n to

TK

E b

alan

ce (

m2 /s

3 )

(b)

Figure 5-3. Analysis of the model representation of the turbulent kinetic energy. (a)Vertical structures of the four terms in the turbulent kinetic energy (TKE)balance (equation (5–8)) for the 10-min measurement burst which started at11:00 UTC on March 31st; (b) TKE balance at 100 cmab; and (c) at 30cmab, vs. time, for all modeled cases. In the order in which they appear inthe right-hand side of equation (5–8), the terms are represented by: circles –shear-induced turbulence production; stars – turbulent diffusion, squares –turbulent dissipation; and “x” – sediment-induced stratification.

82

Page 83: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

0 5 10 150

20

40

60

80

100

120

140

(a)

D

ista

nce

abov

e be

d (c

m)

Suspended sediment concentration (g/L)

0 0.1 0.2 0.3

(b)

North−componentof velocity (m/s)

0 0.2 0.4

(c)

West−componentof velocity (m/s)

10−1

100

101

(d)

Gradient Richardsonnumber

Figure 5-4. Effect of varying floc size D on the model calculations. Vertical structures of:(a) suspended sediment concentration; (b) North- and (c) West-componentsof the mean current; and (d) gradient Richardson number (thick line – 150µm; thin line – 200 µm; dashed line – 300 µm). Observations are marked bycircles. In panel d, the gray line corresponds to the critical Richardsonnumber value of 0.25. The numerical simulations correspond to the 10-minmeasurement burst starting at 11:00 UTC on March 31st (see Figure 5-3a).

83

Page 84: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

CHAPTER 6CONCLUSION

The interaction between bottom turbulence, suspended sediment concentration, and

sediment floc size in cohesive sedimentary environments (e.g., Dyer, 1989; Winterwerp,

1998) was investigated by analyzing field observations and using a bottom boundary

layer model. The data sets were collected during a large scope experiment, conducted

on the muddy Atchafalaya inner shelf, Louisiana, in 2008. During the event which was

focused on in this study, wave energy was moderate (significant height of order of 1 m),

currents were mostly westward and reached speeds of 30 cm/s near bed, and sediment

suspension was relatively dilute with an average concentration of about 2 g/L near bed.

Together with the calculation of wave and current parameters, near-bed Reynolds

stresses were estimated using the differencing-filtering method (Trowbridge, 1998; Shaw

and Trowbridge, 2001; Feddersen and Williams, 2007), which is a challenging task

in wave-energetic environments. In this study, Reynolds stress estimates allowed the

intervals, when bottom turbulence-suspended cohesive sediment interaction was likely

to take place, to be designated, merely by analyzing the observations, even before using

the boundary layer model. These estimates played a key role also in evaluation of the

available information about floc size, as discussed below.

Several elements of the data concur in supporting a weak floc size variability during

the experiment: suspended sediment concentrations show a strong correlation with

short-wave near-bed accelerations, and at the same time, with estimated Reynolds

stresses. A LISST data set, which was collected in the vicinity of the site in 2006,

was evaluated as an independent source of information on suspended sediment size

distribution. Bottom friction velocity was estimated from the fit of logarithmic layers to

the current profile measurements that were co-located with these LISST records. Size

distributions of suspended sediments estimated by LISST, showed remarkable stability

for wave conditions similar to those of the experiment which was focused on here, and

84

Page 85: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

for varying turbulent flow conditions. The major floc mode in the 2006 LISST data was

also consistent with the equilibrium floc sizes (Winterwerp, 1998), calculated using

Reynolds stress estimates and suspended sediment concentration measurements in

2008. It is hard to assess, based on this data set, whether the experimental conditions

were simply not energetic enough to lead to higher concentrations of suspended

sediment, or if a more fundamental physical process might be involved. However, for this

experiment site, and these wave-current conditions, the simplifying assumption that the

floc size is approximately constant seems justified.

While the sediment concentrations recorded were probably not high enough

to modify grain geometry, the values observed are expected to result in flow-field

modifications due to sediment-induced stratification (Winterwerp, 2001, 2006). The

constant floc size assumption allowed the investigation of these effects by using

numerical tools to analyze the observations. Here, a one-dimensional bottom boundary

layer model that was developed for small-scale turbulence and sediment transport

processes on cohesive beds was used (Hsu et al., 2007). The constant floc size input

of the model was obtained from the size distributions in the LISST records. The vertical

structures of flow and suspended sediment concentration were reconstructed. The

model was validated using several test parameters: mean flow, suspended sediment

concentration, and Reynolds stresses. Flows and suspended sediment concentrations

were reproduced accurately; modeled Reynolds stresses matched the estimates based

on observations, with better agreement for cases of strong currents and weak waves,

that were consistent with the low wave-bias formulation of Feddersen and Williams

(2007).

Model simulations show that the turbulence-damping effect of suspended sediment

(i.e., sediment-induced stratification) is important for the turbulence kinetic energy (TKE)

balance throughout the model domain, except in the vicinity of the wave boundary layer,

where shear-induced turbulence production and turbulent dissipation dominate the

85

Page 86: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

balance. In the tidal boundary layer, sediment-induced stratification is of the same order

of magnitude as turbulent dissipation, even without a detectable lutocline (steep gradient

of sediment concentration). In terms of turbulence production, shear-induced turbulence

production appears to dominate over turbulent diffusion in most of the model domain.

The effect of floc size on the numerical results was investigated in the quasi-equilibrium

sense, by carrying out numerical tests with floc size values (within the limits of

observations) under otherwise identical conditions. These tests resulted in virtually

identical Reynolds stress structures. However, smaller floc sizes in the model caused an

increase in the total amount of suspended sediment in the water column (Winterwerp,

2001), that, in turn, resulted in a lower gradient Richardson number increase in the wave

boundary layer.

The results of this study suggest that sediment-induced stratification is an important

mechanism affecting hydrodynamics in muddy environments and therefore has to be

accounted for in practical 3-D circulation models for continental shelf areas. The model

limitations, the dilute concentrations in these observations, i.e., the absence of fluid mud

layers and relatively small contribution of sediment-induced stratification in turbulence

damping near bed, are not allowing an in-depth investigation of mud-induced damping

of surface wave energy. To achieve a better understanding of bottom turbulence -

cohesive sediment interaction and its large scale implications on ocean circulation and

wave dynamics, analysis on a wider range of conditions (more energetic wave action,

lutocline formation, higher suspended sediment concentration values measured) is

essential. The implementation of flocculation models into bottom boundary layer models

(e.g., Winterwerp, 2002) is also likely to improve the modeling capabilities in cohesive

sedimentary environments. Another important contribution to the understanding of

these processes in combined wave-current flow would be to compare model results

with a more detailed vertical structure of suspended sediment concentration and current

(and even Reynolds stresses which could be possible with a current-profiler) near

86

Page 87: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

the bed. This is the topic of an ongoing study in which the backscatter of acoustic

profilers (e.g., ABS and PC-ADP, Figure 2-3) is being calibrated to estimate suspended

sediment concentration throughout the profiling range of the instrumentation (Sahin

et al., 2010). In terms of data analysis methods, direct estimates of the turbulent mass

flux term (φ′w ′), that could be based on the approach of the Reynolds stress estimation

method used herein, would also be useful to evaluate the related closure schemes of the

boundary layer models (equation (5–7)).

The overall findings of this study are summarized in Safak et al. (2010a).

87

Page 88: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

APPENDIX ADIRECT ESTIMATION OF REYNOLDS STRESSES

In this section, derivation of direct estimates of Reynolds stress based on

measurements from one sensor, and from two sensors are detailed. The calculations

are shown for a 2-D coordinate system, let u and w be the horizontal and vertical

components of velocity, respectively. Therefore, the turbulent covariance term of interest

is u′w ′. u and w can be decomposed as follows:

u = u + ~u + u′ , (A–1)

and

w = w + ~w + w ′ , (A–2)

where an overbar denotes a time-averaged quantity, i.e., mean current, a tilde denotes

wave-induced fluctuations, and a prime denotes turbulent fluctuations, respectively. Let

U and W be the horizontal and vertical components of velocity, measured in instrument

coordinates and accounting for their small misalignment θ from the model coordinate

system:

U = (u + u + u′) cos (θ) + (w + w + w ′) sin (θ) ≈ u + u + u′ + θ (w + w + w ′) , (A–3)

and

W = (w + w + w ′) cos (θ)− (u + u + u′) sin (θ) ≈ w + w + w ′ − θ (u + u + u′) . (A–4)

For a single-sensor estimate of Reynolds stress, u′w ′ can be estimated from the

covariance of the horizontal and vertical measurements, i.e., Cov (U, W ) which is equal

to:

88

Page 89: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

Cov(U, W ) =⟨[

U − U] [

W −W]⟩

, (A–5)

where 〈〉 denotes the expected value operator. Substituting demeaned velocities from

equations (A–3) and (A–4) gives:

Cov(U, W ) = 〈[u + u′ + θ (w + w ′)] [w + w ′ − θ (u + u′)]〉 , (A–6)

Cov(U, W ) = 〈uw + uw ′ − θuu − θuu′ + u′w + u′w ′ − θu′u − θu′u′ + θw w + θw w ′

−θ2uw − θ2u′w + θw ′w + θw ′w ′ − θ2uw ′ − θ2u′w ′⟩ . (A–7)

By assuming θ is small, and therefore, by keeping only terms up to order O(θ):

Cov(U, W ) = 〈uw + uw ′ − θuu − θuu′ + u′w + u′w ′ − θu′u − θu′u′ + θw w + θw w ′

+θw ′w + θw ′w ′〉 . (A–8)

A critical step at this point is to assume that wave-induced fluctuations and turbulence-induced

fluctuations are statistically orthogonal (Kitaigorodskii and Lumley, 1983). In other

words, waves and turbulence are assumed to be uncorrelated. Therefore, while

calculating the expected value at the right hand side, terms that include multiplication of

a wave-induced fluctuation and a turbulence-induced fluctuation are assumed to have

zero mean. This gives:

Cov(U, W ) = 〈uw − θuu + u′w ′ − θu′u′ + θw w + θw ′w ′〉 , (A–9)

which can be rewritten as:

Cov(U, W ) = u′w ′ + uw − θ(

uu − w w)− θ

(u′u′ − w ′w ′) . (A–10)

89

Page 90: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

This is the raw estimate of Reynolds stress based on a single sensor’s measurements,

as given in equation (3–24). The first term is the turbulent covariance term of interest.

As explained in Section 3.2.3, assuming that all the components of the turbulent

covariance tensor are of the same order of magnitude (Tennekes and Lumley, 1972), the

last term in (A–10) is O(θ) with respect to the quantity of interest. However, the second

and third terms in equation (A–10) are wave biases and have to be reduced.

Shaw and Trowbridge (2001) presented only the essential steps of the three

versions of the two-sensor method, i.e., by differencing horizontal component, by

differencing vertical component, and by differencing both components. Here, the

two-sensor method that calculates the difference of horizontal components is detailed.

The calculations for the other two versions are similar and the basic steps can be

followed from Shaw and Trowbridge (2001). In the 2-D coordinate system presented

above, the horizontal and vertical velocity measurements from two sensors can be

written as:

U1 ≈ u1 + w1θ1 ; W1 ≈ w1 − u1θ1 , (A–11)

U2 ≈ u2 + w2θ2 ; W2 ≈ w2 − u2θ2 . (A–12)

Subscript “1” is set to refer to the sensor at which the Reynolds stress is estimated.

Since waves and turbulence are assumed to be uncorrelated (Kitaigorodskii and Lumley,

1983), the covariance of the horizontal velocity difference between the two sensors,

and the vertical velocity at the sensor of interest, can be decomposed into its wave and

turbulence components:

Cov(�U, W1) = Cov(� ~U, ~W1) + Cov(�U ′, W ′1) , (A–13)

90

Page 91: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

where � denotes differencing operator (herein, subtraction of velocity measurements at

sensor 2 from those at sensor 1). The turbulence component is given as:

Cov(�U ′, W ′1) =

⟨[�U ′ − �U ′] [

W ′1 −W ′

1

]⟩, (A–14)

Cov(�U ′, W ′1) =

⟨[U ′

1 − U ′2 − U ′

1 − U ′2

] [W ′

1 −W ′1

]⟩. (A–15)

By definition, turbulence-induced fluctuations have a zero mean, which simplifies (A–15)

as:

Cov(�U ′, W ′1) = 〈[U ′

1 − U ′2] W ′

1〉 , (A–16)

Cov(�U ′, W ′1) = 〈[u′1 + w ′

1θ1 − u′2 − w ′2θ2] [w ′

1 − u′1θ1]〉 , (A–17)

Cov(�U ′, W ′1) =

⟨u′1w ′

1 − θ1u′1u′1 + θ1w ′1w ′

1 − θ21u′1w ′

1 − u′2w ′1 + θ1u′1u′2

−θ2w ′1w ′

2 + θ1θ2u′1w ′2〉 . (A–18)

By keeping only terms upto order O(θ):

Cov(�U ′, W ′1) = 〈u′1w ′

1 − θ1 (u′1u′1 − w ′1w ′

1)− u′2w ′1 + θ1u′1u′2 − θ2w ′

1w ′2〉 , (A–19)

which can be rewritten as:

Cov(�U ′, W ′1) = u′1w ′

1 − θ1(

u′1u′1 − w ′1w ′

1

)− u′2w ′1 + u′1u′2θ1 − w ′

1w ′2θ2 . (A–20)

The first term on the right-hand side is the quantity of interest. The second term is an

order of magnitude smaller than the quantity of interest (Tennekes and Lumley, 1972).

91

Page 92: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

The last three terms describe cross-sensor turbulence correlations, which bias the

Reynolds stress estimates. The most significant of these terms is u′2w ′1, because the

other two terms are O(θ). In order to reduce the effect of this turbulence bias term, the

distances between the measurement points and the bed need to be adjusted (Section

3.2.3). Assuming that the experimental setup is prepared properly, Cov(�U ′, W ′1) yields

an estimate of the quantity of interest, with reduced wave bias compared to the single

sensor estimate (Equation (A–10)).

Wave component in equation (A–13) is calculated as:

Cov(� ~U, ~W1) =⟨[

� ~U − � ~U] [

~W1 − ~W1

]⟩, (A–21)

Cov(� ~U, ~W1) =⟨[

~U1 − ~U2 − ~U1 − ~U2

] [~W1 − ~W1

]⟩. (A–22)

By definition, wave-induced fluctuations have a zero mean, this yields:

Cov(� ~U, ~W1) =⟨[

~U1 − ~U2

]~W1

⟩. (A–23)

Velocities in instrument coordinates are substituted through equations (A–11) and

(A–12):

Cov(� ~U, W1) = 〈[~u1 + ~w1θ1 − ~u2 − ~w2θ2] [ ~w1 − ~u1θ1]〉 , (A–24)

Cov(� ~U, W1) =⟨

~u1 ~w1 − θ1~u1~u1 + θ1 ~w1 ~w1 − θ21~u1 ~w1 − ~u2 ~w1 + θ1~u1~u2 − θ2 ~w1 ~w2 + θ1θ2~u1 ~w2

⟩.

(A–25)

Keeping only the terms up to order O(θ) gives:

Cov(� ~U, W1) = 〈~u1 ~w1 − θ1~u1~u1 + θ1 ~w1 ~w1 − ~u2 ~w1 + θ1~u1~u2 − θ2 ~w1 ~w2〉 , (A–26)

92

Page 93: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

which can be compiled as:

Cov(� ~U, ~W1) = ~w1�~u + θ1( ~w1� ~w − ~u1�~u) + �θ( ~w 21 − ~w1� ~w ) . (A–27)

At a stationary point in the horizontal plane, a sinusoidal wave of height H and

radian frequency ω at a water depth h induce the following horizontal and vertical orbital

velocities:

~u(t) =Hω

2cosh(kw z)sinh(kw h)

cos (ωt) , (A–28)

~w (t) = −Hω

2sinh(kw z)sinh(kw h)

sin (ωt) , (A–29)

where z is the height above the bed, and kw is the wavenumber corresponding to ω at

depth h. Vertical gradients of these orbital velocities are calculated as:

∂~u(t)∂z

=Hωkw

2sinh(kw z)sinh(kw h)

cos(ωt) = ~u(t)kw tanh(kw z) , (A–30)

∂ ~w (t)∂z

= −Hωkw

2cosh(kw z)sinh(kw h)

sin (ωt) =~w (t).kw

tanh(kw z). (A–31)

Vertical gradients are represented as first order expansions (Shaw and Trowbridge,

2001) for vertical separation �z = r :

�~u = r ~ukw tanh(kw z) ; � ~w =r ~w kw

tanh(kw z). (A–32)

Substituting these terms into equation (A–27) gives:

Cov(� ~U, ~W1) = ~w1r ~u1kw tanh(kw z) + θ1

(~w1

r ~w1kw

tanh(kw z)− ~u1r ~u1kw tanh(kw z)

)

+�θ

(~w 2

1 − ~w1r ~w1kw

tanh(kw z)

), (A–33)

93

Page 94: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

Cov(� ~U, ~W1) = rkw tanh(kw z) ~u1 ~w1 + θ1

(~w 2

1rkw

tanh(kw z)− rkw tanh(kw z)~u2

1

)

+�θ ~w 21

(1− rkw

tanh(kw z)

), (A–34)

Cov(� ~U, ~W1) =rz

[kw ztanh(kw z) ~u1 ~w1 + θ1

zr

(~w 2

1rkw

tanh(kw z)− rkw tanh(kw z)~u2

1

)

+�θ ~w 21

zr

(1− rkw

tanh(kw z)

)]. (A–35)

Near the bottom, i.e., for small z , tanh(kw z) ≈ kw z . This yields:

Cov(� ~U, ~W1) =rz

[k2

w z 2 ~u1 ~w1 + θ1 ~w 21 − k2

w z 2θ1~u21 + �θ ~w 2

1zr− �θ ~w 2

1

]. (A–36)

Near the sea bed, the first and third terms on the right hand side are expected to be

larger than the other three terms. Therefore, the wave component of the Reynolds stress

estimate can be approximated as:

Cov(� ~U, ~W1) ≈ rk2w z

[~u1 ~w1 − θ1~u2

1

]. (A–37)

Applying the differencing to the horizontal velocity components reduces the wave bias in

the one-sensor estimate (equation (A–10)) with a factor of rk2w z .

94

Page 95: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

REFERENCES

Agrawal, Y. C., Aubrey, D. G., 1992. Velocity observations above a rippled bed usinglaser doppler velocimetry. J. Geophys. Res. 97 (C12), 20249–20259.

Agrawal, Y. C., Pottsmith, H. C., 1994. Laser diffraction particle sizing in STRESS. Cont.Shelf Res. 14, 1101–1121.

Allison, M. A., Kineke, G. C., Gordon, E. S., Goni, M. A., 2000. Development andreworking of a seasonal flood deposit on the inner continental shelf off the AtchafalayaRiver. Cont. Shelf Res. 20, 2267–2294.

Allison, M. A., Sheremet, A., Goni, M. A., Stone, G. W., 2005. Storm layer deposition onthe Mississippi-Atchafalaya subaqueous delta generated by Hurricane Lili in 2002.Cont. Shelf Res. 25, 2213–2232.

Allison, M. A., Sheremet, A., Safak, I., Duncan, D. D., 2010. Floc behavior in highturbidity coastal settings as recorded by LISST: the Atchafalaya Delta inner shelf,Louisiana. AGU Ocean Sciences’10 Meeting , Portland, Oregon. Poster presentation.

Bendat, J. S., Piersol, A. G., 1971. Random Data: Analysis and MeasurementProcedures. Wiley-Interscience.

Benilov, A. Y., Filyushkin, B. N., 1970. Application of methods of linear filtration to ananalysis of fluctuations in the surface layer of the sea. Izv. Atmos. Oceanic Phys. 6,810–819.

Berhane, I., Sternberg, R. W., Kineke, G. C., Milligan, T. G., Kranck, K., 1997. Thevariability of suspended aggregates on the Amazon Continental Shelf. Cont. ShelfRes. 17, 267–285.

Bricker, J. D., Monismith, S. G., 2007. Spectral wave-turbulence decomposition. J.Atmos. Ocean. Tech. 24, 1479–1487.

Cacchione, D. A., Drake, D. E., Kayen, R. W., Sternberg, R. W., Kineke, G. C., Tate,G. B., 1995. Measurements in the bottom boundary layer on the Amazon subaqueousdelta. Mar. Geo. 125, 235–257.

Dean, R. G., Dalrymple, R. A., 1991. Water Wave Mechanics for Engineers andScientists. World Scientific, Adv. Series on Ocean Eng. 2.

Di Marco, S. F., Reid, R. O., 1998. Characterization of the principal tidal currentconstituents on the Texas-Louisiana shelf. J. Geophys. Res. 103, 3093–3109.

Downing, J. P., Sternberg, R. W., Lister, C. R. B., 1981. New instrumentation for theinvestigation of sediment suspension processes in the shallow marine environment.Mar. Geo. 42, 19–34.

95

Page 96: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

Draut, A. E., Kineke, G. C., Velasco, D. W., Allison, M. A., Prime, R. J., 2005. Influence ofthe Atchafalaya River on recent evolution of the chenier-plain inner continental shelf,northern Gulf of Mexico. Cont. Shelf Res. 25, 91–112.

Dyer, K. R., 1989. Sediment processes in estuaries: Future research requirements. J.Geophys. Res. 97, 14327–14339.

Elgar, S., Raubenheimer, B., 2008. Wave dissipation by muddy seafloors. Geophys. Res.Let. 35, L07611, doi:10.1029/2008GL033245.

Elgar, S., Raubenheimer, B., Guza, R. T., 2005. Quality control of acoustic Dopplervelocimeter data in the surf zone. Meas. Sci. Tech. 16, 1889–1893.

Feddersen, F., Williams, A. J., 2007. Direct estimation of the Reynolds stress verticalstructure in the nearshore. J. Atmos. Ocean. Tech. 24, 102–116.

Fredsoe, J., Deigaard, R., 1992. Mechanics of Coastal Sediment Transport. WorldScientific, Adv. Series on Ocean Eng. 3.

Fugate, D. C., Friedrichs, C. T., 2002. Determining concentration and fall velocity ofestuarine particle populations using ADV, OBS and LISST. Cont. Shelf Res. 22,1867–1886.

Goni, M. A., Gordon, E. S., Monacci, N. M., Clinton, R., Gisewhite, R., Allison, M. A.,Kineke, G. C., 2006. The effect of Hurricane Lili on the distribution of organicmatter along the inner Louisiana shelf (Gulf of Mexico, USA). Cont. Shelf Res. 26,2260–2280.

Grant, H. L., 1958. The large eddies of turbulent motion. J. Fluid Mech. 4, 149–190.

Grant, W. D., Madsen, O. S., 1979. Combined wave and current interaction with a roughbottom. J. Geophys. Res. 84, 1797–1808.

Grant, W. D., Madsen, O. S., 1986. The continental-shelf bottom boundary layer. Ann.Rev. Fluid Mech. 18, 265–305.

Gratiot, N., Michallet, H., Mory, M., 2005. On the determination of the settlingflux of cohesive sediments in a turbulent fluid. J. Geophys. Res. 110 (C06004),doi:10.1029/2004JC002732.

Gratiot, N., Mory, M., Auchere, D., 2000. An acoustic doppler velocimeter (ADV) forthe characterisation of turbulence in concentrated fluid mud. Cont. Shelf Res. 20,1551–1567.

Hasselmann, D. E., Dunckel, M., Ewing, J. A., 1980. Directional wave spectra observedduring JONSWAP 1973. J. Phys. Oceanogr. 10, 1264–1280.

96

Page 97: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

Herbers, T. H. C., Guza, R. T., 1993. Comment on ‘Velocity observations above a rippledbed using laser doppler velocimetry’ by Y.C. Agrawal and D.G. Aubrey. J. Geophys.Res. 98, 20331–20333.

Hill, P. S., 1998. Controls on floc size in the sea. Oceanography 11, 13–18.

Hoefel, F., Elgar, S., 2003. Wave-induced sediment transport and sandbar migration.Science 299, 1885–1887.

Hsu, T.-J., Hanes, D. M., 2004. Effects of wave shape on sheet flow sediment transport.J. Geophys. Res. 109 (C05025), doi:10.1029/2003JC002075.

Hsu, T.-J., Liu, P. L.-F., 2004. Toward modeling turbulent suspension of sand in thenearshore. J. Geophys. Res. 109 (C05025), doi:10.1029/2003JC002240.

Hsu, T.-J., Ozdemir, C. E., Traykovski, P. A., 2009. High-resolution numerical modelingof wave-supported gravity-driven mudflows. J. Geophys. Res. 114 (C05014),doi:10.1029/2008JC005006.

Hsu, T.-J., Traykovski, P. A., Kineke, G. C., 2007. On modeling boundary layerand gravity-driven fluid mud transport. J. Geophys. Res. 112 (C04011),doi:10.1029/2006JC003719.

Jaramillo, S., 2008. Observations of wave-sediment interactions on a muddy shelf. Ph.D. Thesis , University of Florida, Gainesville.

Jaramillo, S., Sheremet, A., Allison, M. A., Reed, A. T., Holland, K. T., 2009.Wave-mud interactions over the muddy Atchafalaya subaqueous clinoform,Louisiana, USA: Wave-driven sediment transport. J. Geophys. Res. 114 (C04002),doi:10.1029/2008JC004821.

Jiang, F., Mehta, A. J., 1996. Mudbanks of the southwest coast of india. V: Waveattenuation. J. Coas. Res. 12, 890–897.

Jiang, F., Mehta, A. J., 2000. Lutocline behavior in high-concentration estuary. J. Wtrwy.,Port, Coast. and Ocean Eng., ASCE, 126, 324–328.

Jonsson, I. G., 1966. Wave boundary layers. Proc. 10th Conf. Coas. Eng, ASCE, Tokyo,127–148.

Kaihatu, J. M., Sheremet, A., Holland, K. T., 2007. A model for the propagation ofnonlinear surface waves over viscous muds. Coas. Eng. 54, 752–764.

Kaimal, J. C., Wyngaard, J. C., Izumi, Y., Cote, O. R., 1972. Spectral characteristics ofsurface layer turbulence. Quart. J. Roy. Meteor. Soc. 98, 563–589.

Kamphuis, J. W., 1975. Friction factor under oscillatory waves. J. Wtrwy., Port, Coast.Eng. Div., ASCE, 101, 135–144.

97

Page 98: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

Khelifa, A., Hill, P. S., 2006. Models for effective density and settling velocity of flocs. J.Hydraul. Res. 44, 390–401.

Kineke, G. C., Higgins, E. E., Hart, K., Velasco, D. W., 2006. Fine-sediment transportassociated with cold-front passages on the shallow shelf, Gulf of Mexico. Cont. ShelfRes. 26, 2073–2091.

Kitaigorodskii, S. A., Lumley, J. L., 1983. Wave-turbulence interactions in the upperocean. Part 1: The energy balance of the interacting fields of surface wind waves andwind-induced three-dimensional turbulence. J. Phys. Oceanogr. 13, 11977–11987.

Kranenburg, C., 1994. The fractal structure of cohesive sediment aggregates. Est. Coas.and Shelf Sci. 39, 451–460.

Lacy, J. R., Sherwood, C. R., Wilson, D. J., Chisholm, T. A., Gelfenbaum, G. R.,2005. Estimating hydrodynamic roughness in a wave-dominated environmentwith a high-resolution acoustic Doppler profiler. J. Geophys. Res. 110 (C06014),doi:10.1029/2003JC001814.

Levich, V. G., 1962. Physicochemical Hydrodynamics. Prentice-Hall, Inc.

Lhermitte, R., Serafin, R. J., 1984. Pulse-to-pulse coherent Doppler sonar signalprocessing techniques. J. Atmos. Ocean. Tech. 1, 293–308.

Lohrmann, A., Hackett, B., Roed, L. P., 1990. High resolution measurements ofturbulence, velocity and stress using a pulse-to-pulse coherent sonar. J. Atmos.Ocean. Tech. 7, 19–37.

Lueck, R., Lu, Y., 1997. The logarithmic layer in a tidal channel. Cont. Shelf Res. 17,1785–1801.

Mehta, A. J., 1989. On estuarine cohesive sediment suspension behaviour. J. Geophys.Res. 94, 14303–14314.

Mehta, A. J., 2002. Mudshore dynamics and controls. Muddy coasts of the world:Processes, deposits and function , T.Healy, Y. Wang and J. A. Healy eds., Elsevier,Amsterdam, 19–60.

Michallet, H., Mory, M., 2004. Modeling sediment suspensions in oscillating gridturbulence. Fluid Dyn. Res. 35, 87–106.

Monismith, S. G., Magnaudet, J., 1998. On wavy mean flows, Langmuir cells, strain,and turbulence. Physical Processes in Lakes and Oceans , J.Imberger, ed., AmericanGeophysical Union, 101–110.

Mossa, J., 1996. Sediment dynamics in the lowermost Mississippi River. Eng. Geo. 45,457–479.

98

Page 99: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

Neill, C. F., Allison, M. A., 2005. Subaqueous deltaic formation on the Atchafalaya Shelf,Louisiana. Mar. Geo. 214, 411–430.

Ng, C.-N., 2000. Water waves over a muddy bed: a two-layer Stokes’ boundary layermodel. Coas. Eng. 40, 221–242.

Nielsen, P., 1992. Coastal Bottom Boundary Layers and Sediment Transport. WorldScientific, Adv. Series on Ocean Eng. 4.

Parker, W. R., Kirby, R., 1982. Time dependent properties of cohesive sediment relevantto sedimentation management-A European experience. Estuarine comparisons , V.S.Kennedy, ed., Academic, 573–590.

Phillips, O. M., 1958. The equilibrium range in the spectrum of wind-generated waves. J.Fluid Mech. 4, 426–434.

Priestley, M. B., 1981. Spectral Analysis and Time Series. Academic Press.

Rosman, J. H., Hench, J. L., Koseff, J. R., Monismith, S. G., 2008. Extracting Reynoldsstresses from acoustic doppler current profiler measurements in wave-dominatedenvironments. J. Atmos. Ocean. Tech. 25, 286–306.

Safak, I., Sheremet, A., Allison, M. A., Hsu, T.-J., 2010a. Bottom turbulence on themuddy Atchafalaya Shelf, Louisiana, USA. under review in J. Geophys. Res.

Safak, I., Sheremet, A., Elgar, S., Raubenheimer, B., 2010b. Nonlinear wavepropagation across a muddy seafloor. AGU Ocean Sciences’10 Meeting , Portland,Oregon. Oral presentation.

Sahin, C., Safak, I., Sheremet, A., Allison, M. A., 2010. A method for estimatingconcentration profiles for suspended cohesive sediment based on profiles ofacoustic backscatter. AGU Ocean Sciences’10 Meeting , Portland, Oregon. Posterpresentation.

Scott, C. P., Cox, D. T., Maddux, T. B., Long, J. W., 2005. Large-scale laboratoryobservations of turbulence on a fixed barred beach. Meas. Sci. Tech. 16, 1903–1912.

Shaw, W. J., Trowbridge, J. H., 2001. The direct estimation of near-bottom turbulentfluxes in the presence of energetic wave motions. J. Atmos. Ocean. Tech. 18,1540–1557.

Sheremet, A., Jaramillo, S., Su, S.-F., Allison, M. A., Holland, K. T., 2010. Wave-mudinteractions over the muddy Atchafalaya subaqueous clinoform, Louisiana, Usa:Mud-induced wave dissipation. under review in J. Geophys. Res.

Sheremet, A., Mehta, A. J., Liu, B., Stone, G. W., 2005. Wave-sediment interaction on amuddy inner shelf during Hurricane Claudette. Est. Coas. and Shelf Sci. 63, 225–233.

99

Page 100: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

Sheremet, A., Stone, G. W., 2003. Observations of nearshore wave dissipation overmuddy sea beds. J. Geophys. Res. 108 (C11,3357), doi:10.1029/2003JC001885.

Son, M., Hsu, T.-J., 2009. The effect of variable yield strength and variable fractaldimensional on flocculation of cohesive sediment. Water Res. 43, 3582–3592.

SonTek/YSI, 2001. Acoustic Doppler Velocimeter-Hydra technical documentation. 6837Nancy Ridge Drive, San Diego, CA 92121.

Soulsby, R. L., 1977. Similarity scaling of turbulence spectra in marine and atmosphericboundary layers. J. Phys. Oceanogr. 7, 934–937.

Soulsby, R. L., 1980. Selecting record length and digitization rate for near-bedturbulence measurements. J. Phys. Oceanogr. 10, 208–219.

Stacey, M. T., Monismith, S. G., Burau, J. R., 1999. Measurements of Reynolds stressprofiles in an unstratified tidal flow. J. Geophys. Res. 104 (C5), 10933–10949.

Sterling, G. H., Strohbeck, E. E., 1973. The failure of the South Pass 70 B Platform inHurricane Camille. 5th Ann. Offshore Tech. Conf., Houston, TX, 719–724.

Stevens, A. W., Wheatcroft, R. A., Wiberg, P. L., 2007. Seabed properties and sedimenterodibility along the western Adriatic margin, Italy. Cont. Shelf Res. 27, 400–416.

Styles, R. B., 1998. A continental shelf bottom boundary layer model: Development,calibration and applications to sediment transport in the Middle Atlantic Bight. Ph. D.Thesis , Rutgers, the State University of New Jersey.

Styles, R. B., Glenn, S. M., 2000. Modeling stratified wave and current bottom boundarylayers on the continental shelf. J. Geophys. Res. 105 (C10), 24119–24139.

Styles, R. B., Glenn, S. M., 2002. Modeling bottom roughness in thepresence of wave-generated ripples. J. Geophys. Res. 107 (C8,3110),doi:10.1029/2001JC000864.

Tennekes, H., Lumley, J. L., 1972. A First Course in Turbulence. The MIT Press.

Traykovski, P., Geyer, W. R., Irish, J. D., Lynch, J. F., 2000. The role of wave-induceddensity-driven fluid mud flows for cross-shelf transport on the Eel River continentalshelf. Cont. Shelf Res. 20, 2113–2140.

Traykovski, P., Wiberg, P. L., Geyer, W. R., 2007. Observations and modeling ofwave-supported sediment gravity flows on the Po prodelta and comparison to priorobservations from the Eel shelf. Cont. Shelf Res. 27, 375–399.

Trowbridge, J. H., 1998. On a technique for measurement of turbulent shear stress in thepresence of surface waves. J. Atmos. Ocean. Tech. 15, 290–298.

Trowbridge, J. H., Agrawal, Y. C., 1995. Glimpses of a wave boundary layer. J. Geophys.Res. 100 (C10), 20729–20743.

100

Page 101: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

Trowbridge, J. H., Elgar, S., 2001. Turbulence measurements in the surf zone. J. Phys.Oceanogr. 31, 2403–2417.

Trowbridge, J. H., Elgar, S., 2003. Spatial scales of stress-carrying nearshoreturbulence. J. Phys. Oceanogr. 33, 1122–1128.

Trowbridge, J. H., Geyer, W. R., Bowen, M. M., Williams III, A. J., 1999. Near-bottomturbulence measurements in a partially mixed estuary: Turbulent energy balance,velocity structure, and along-channel momentum balance. J. Phys. Oceanogr. 29,3056–3072.

Trowbridge, J. H., Kineke, G. C., 1994. Structure and dynamics of fluid muds on theAmazon continental shelf. J. Geophys. Res. 99 (C1), 865–874.

Vinzon, S. B., Mehta, A. J., 1998. Mechanism for formation of lutoclines by waves. J.Wtrwy., Port, Coast. and Ocean Eng., ASCE, 124, 147–149.

Voulgaris, G., Trowbridge, J. H., 1998. Evaluation of the acoustic dopplervelocimeter(ADV) for turbulence measurements. J. Atmos. Ocean. Tech. 15, 272–289.

Walker, N. D., Hammack, A. B., 2000. Impacts of winter storms on circulation andsediment transport: Atchafalaya-Vermilion Bay region, Louisiana, USA. J. Coas. Res.16, 996–1010.

Wells, J. T., Kemp, G. P., 1981. Atchafalaya mud stream and recent mudflatprogradation: Louisiana chenier plain. Gulf Coast Assoc. Geol. Soc. Trans. 31,409–416.

Whipple, A. C., Luettich Jr., R. A., Seim, H. E., 2006. Measurements of Reynolds stressin a wind-driven lagoonal estuary. Ocean Dyn. 56, 169–185.

Winterwerp, J. C., 1998. A simple model for turbulence induced flocculation of cohesivesediment. J. Hydraul. Res. 36, 309–326.

Winterwerp, J. C., 2001. Stratification effects by cohesive and non-cohesive sediment. J.Geophys. Res. 106, 22559–22574.

Winterwerp, J. C., 2002. On the flocculation and settling velocity of estuarine mud. Cont.Shelf Res. 22, 1339–1360.

Winterwerp, J. C., 2006. Stratification effects by fine suspended sediment atlow, medium, and very high concentrations. J. Geophys. Res. 111 (C05012),doi:10.1029/2005JC003019.

Winterwerp, J. C., de Graaff, R. F., Groeneweg, J., Luijendijk, A. P., 2007. Modelling ofwave damping at Guyana mud coast. Coas. Eng. 54, 249–261.

Winterwerp, J. C., van Kesteren, W. G. M., 2004. Introduction to the Physics of CohesiveSediment in the Marine Environment. Elsevier.

101

Page 102: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

Wolf, J., 1999. The estimation of shear stresses from near-bed turbulent velocities forcombined wave-current flows. Coas. Eng. 37, 529–543.

Wright, L. D., Friedrichs, C. T., Kim, S. C., Scully, M. E., 2001. Effects of ambientcurrents and waves on gravity-driven sediment transport on continental shelves. Mar.Geo. 175, 25–45.

Zarraga, I. E., Hill, D. A., Leighton, D. T., 2000. The characterization of the total stressof concentrated suspension of noncolloidal spheres in newtonian fluids. J. Rheol. 44,185–220.

102

Page 103: INTERACTION OF BOTTOM TURBULENCE AND COHESIVE SEDIMENT … · interaction of bottom turbulence and cohesive sediment on the muddy atchafalaya shelf, louisiana, usa by ... 1 introduction

BIOGRAPHICAL SKETCH

Ilgar Safak was born in 1982 in Ankara, Turkey. He graduated from T.E.D. Ankara

College in 1999. He earned his B.S. in Civil Engineering in 2003, and M.S. in Coastal

Engineering in 2006, both from Middle East Technical University in Ankara. During his

M.S. studies, he worked on physical modeling of performance of floating breakwaters,

and numerical modeling of wind-wave induced longshore sediment transport. Since

August 2006, he has been working with Alex Sheremet at the University of Florida.

His research project is on the interaction of hydrodynamics and sediment transport

processes in wave-energetic muddy environments. After he earns his Ph.D. degree, he

would like to enrich his understanding of coastal and oceanographic processes.

103