solute dispersion in the coastal boundary layer of ...phani/jgrc20136.pdf · solute dispersion in...

12
Solute dispersion in the coastal boundary layer of southern Lake Michigan Pramod Thupaki, 1 Mantha S. Phanikumar, 1 and Richard L. Whitman 2 Received 15 June 2012; revised 5 February 2013; accepted 15 February 2013; published 29 March 2013. [1] We evaluate a three-dimensional, nested-grid nearshore model of Lake Michigan for its ability to describe key aspects of hydrodynamics and solute transport using data from a eld study conducted in summer 2008. Velocity comparisons with observations from ve bottom-mounted ADCPs at different depths in the coastal boundary layer (CBL) show that the numerical model was able to simulate currents and ow reversals accurately within the inertial boundary layer, however model accuracy reduced close to the shoreline. Power spectra of observed and simulated velocity time series at different locations showed that the hydrodynamic model was able to describe the energy contained in the inertial scales, but over-predicted turbulent dissipation rates. As a result, model-predicted values of energy in the smaller, dissipation scales were lower compared to values calculated from observations in the CBL. Differences between energy contained in the observed and simulated velocity spectra increased as the shoreline is approached. Observations showed that vertical variations in the alongshore and cross-shore velocities were dominated by inertial waves. Inaccuracies in representing energy dissipation rates and processes not explicitly described in the hydrodynamic model (e.g., anisotropy, waves) could potentially contribute to errors in describing transport in the CBL. Measurements from a continuous dye release experiment from a riverine outfall were described using a nearshore model with a mean horizontal dispersion coefcient of 5.6 m 2 /s. Improved representations of physical processes (such as turbulence, internal waves and wave-current interactions) can be expected to provide better descriptions of solute transport in the CBL. Citation: Thupaki, P., M. S. Phanikumar, and R. L. Whitman (2013), Solute dispersion in the coastal boundary layer of southern Lake Michigan, J. Geophys. Res. Oceans, 118, 1606–1617, doi:10.1002/jgrc.20136. 1. Introduction [2] Addressing science questions aimed at effective management of coastal resources often calls for the ability to accurately describe the dispersion of material in the nearshore region. Questions involving beach closures and the fate and transport of pathogenic microorganisms such as bacteria, viruses and protozoan parasites are intimately linked to solute dispersion. Recreational beaches in southern Lake Michigan are impacted by microbial contamination entering the nearshore environment via discharges from riverine outfalls in addition to contributions from non-point sources. One of the major sources of contamination at river sites is discharge from combined sewer overows (CSOs). Beaches are closed to the public whenever levels of fecal indicator bacteria (FIB) such as Escherichia coli exceed local standards, with signicant economic and health-risk tradeoffs associated with beach closures [Rabinovici et al., 2004; Dorfman and Rosselot, 2010]. Traditional methods of beach management are observation-based and require 24 hours to run the assays during which time conditions at the beaches can change. Hence, well-tested hydrodynamic and transport models, with the potential to make short-term forecasts, are attractive tools for beach management. [3] A number of biotic and abiotic factors inuence the transport, inactivation (loss per unit time) and survival of FIB in the nearshore environment. Key processes include bacterial inactivation due to sunlight (direct DNA damage) with turbid waters contributing to increased survival by limiting sunlight penetration, removal of bacteria from the water column due to sedimentation after attachment to suspended particles, increased loss of bacteria at elevated temperatures (due to damage of bacterial cell components as well as increased predation risk), and dependence on pH, nutrient content and dissolved oxygen, among other factors [Hipsey et al., 2008]. Although the relative importance of these factors can change depending on the geographical setting, beach orientation and beach type (open versus embayed), nearshore hydrodynamic and wave processes play a key role in controlling overall FIB fate and transport [Ge et al., 2012a, 2012b]. A budget analysis including 1 Department of Civil and Environmental Engineering, Michigan State University, East Lansing, Michigan, USA. 2 United States Geological Survey, Great Lakes Science Center, Porter, Indiana, USA. Corresponding author: M. S. Phanikumar, Department of Civil and Environmental Engineering, 1449 Engineering Research Court, Room A130, Michigan State University, East Lansing, MI 48824, USA. ([email protected]) ©2013. American Geophysical Union. All Rights Reserved. 2169-9275/13/10.1002/jgrc.20136 1606 JOURNAL OF GEOPHYSICAL RESEARCH: OCEANS, VOL. 118, 16061617, doi:10.1002/jgrc.20136, 2013

Upload: ledat

Post on 31-Aug-2018

214 views

Category:

Documents


0 download

TRANSCRIPT

Solute dispersion in the coastal boundary layer of southernLake Michigan

Pramod Thupaki,1 Mantha S. Phanikumar,1 and Richard L. Whitman2

Received 15 June 2012; revised 5 February 2013; accepted 15 February 2013; published 29 March 2013.

[1] We evaluate a three-dimensional, nested-grid nearshore model of Lake Michigan forits ability to describe key aspects of hydrodynamics and solute transport using data from afield study conducted in summer 2008. Velocity comparisons with observations from fivebottom-mounted ADCPs at different depths in the coastal boundary layer (CBL) show thatthe numerical model was able to simulate currents and flow reversals accurately within theinertial boundary layer, however model accuracy reduced close to the shoreline. Powerspectra of observed and simulated velocity time series at different locations showed that thehydrodynamic model was able to describe the energy contained in the inertial scales, butover-predicted turbulent dissipation rates. As a result, model-predicted values of energy inthe smaller, dissipation scales were lower compared to values calculated from observationsin the CBL. Differences between energy contained in the observed and simulated velocityspectra increased as the shoreline is approached. Observations showed that verticalvariations in the alongshore and cross-shore velocities were dominated by inertial waves.Inaccuracies in representing energy dissipation rates and processes not explicitly describedin the hydrodynamic model (e.g., anisotropy, waves) could potentially contribute to errorsin describing transport in the CBL. Measurements from a continuous dye releaseexperiment from a riverine outfall were described using a nearshore model with a meanhorizontal dispersion coefficient of 5.6 m2/s. Improved representations of physicalprocesses (such as turbulence, internal waves and wave-current interactions) can beexpected to provide better descriptions of solute transport in the CBL.

Citation: Thupaki, P., M. S. Phanikumar, and R. L. Whitman (2013), Solute dispersion in the coastal boundary layer ofsouthern Lake Michigan, J. Geophys. Res. Oceans, 118, 1606–1617, doi:10.1002/jgrc.20136.

1. Introduction

[2] Addressing science questions aimed at effectivemanagement of coastal resources often calls for the abilityto accurately describe the dispersion of material in thenearshore region. Questions involving beach closures andthe fate and transport of pathogenic microorganisms suchas bacteria, viruses and protozoan parasites are intimatelylinked to solute dispersion. Recreational beaches in southernLake Michigan are impacted by microbial contaminationentering the nearshore environment via discharges fromriverine outfalls in addition to contributions from non-pointsources. One of the major sources of contamination at riversites is discharge from combined sewer overflows (CSOs).Beaches are closed to the public whenever levels of fecalindicator bacteria (FIB) such as Escherichia coli exceed

local standards, with significant economic and health-risktradeoffs associated with beach closures [Rabinoviciet al., 2004; Dorfman and Rosselot, 2010]. Traditionalmethods of beach management are observation-basedand require 24 hours to run the assays during which timeconditions at the beaches can change. Hence, well-testedhydrodynamic and transport models, with the potentialto make short-term forecasts, are attractive tools forbeach management.[3] A number of biotic and abiotic factors influence the

transport, inactivation (loss per unit time) and survival ofFIB in the nearshore environment. Key processes includebacterial inactivation due to sunlight (direct DNA damage)with turbid waters contributing to increased survival bylimiting sunlight penetration, removal of bacteria from thewater column due to sedimentation after attachment tosuspended particles, increased loss of bacteria at elevatedtemperatures (due to damage of bacterial cell componentsas well as increased predation risk), and dependence onpH, nutrient content and dissolved oxygen, among otherfactors [Hipsey et al., 2008]. Although the relative importanceof these factors can change depending on the geographicalsetting, beach orientation and beach type (open versusembayed), nearshore hydrodynamic and wave processesplay a key role in controlling overall FIB fate and transport[Ge et al., 2012a, 2012b]. A budget analysis including

1Department of Civil and Environmental Engineering, Michigan StateUniversity, East Lansing, Michigan, USA.

2United States Geological Survey, Great Lakes Science Center, Porter,Indiana, USA.

Corresponding author: M. S. Phanikumar, Department of Civil andEnvironmental Engineering, 1449 Engineering Research Court, Room A130,Michigan State University, East Lansing, MI 48824, USA. ([email protected])

©2013. American Geophysical Union. All Rights Reserved.2169-9275/13/10.1002/jgrc.20136

1606

JOURNAL OF GEOPHYSICAL RESEARCH: OCEANS, VOL. 118, 1606–1617, doi:10.1002/jgrc.20136, 2013

physical and biological processes affecting FIB fate andtransport indicated that dilution of FIB due to mixing is akey process that dominates overall transport [Thupaki et al.,2010]. Since peak concentrations of bacteria are determinedby the combined action of dilution and other loss processes,dilution rates should be accurately represented to improvemodel performance. Therefore, the ability to describe trans-port of a conservative tracer accurately is a necessary first stepto modeling bacterial fate and transport.[4] Small-scale features dominate the mixing and transport

of riverine plumes in the nearshore coastal boundary layer.Analysis of river plume mixing dynamics shows that thespatial domain can be separated into: (a) the near-field regionwhere mixing is dominated by buoyancy and momentum ofthe plume and (b) the far-field region where plume dynamicsis primarily controlled by turbulent diffusion. Based on fieldobservations of the Grand River plume in Lake Michigan,[Nekouee, 2010] classified surface buoyant plumes intotwo major categories - shore-attached plumes, which occurwhen along-shore currents are strong, and unattachedplumes. Unattached plumes were further classified into fivesub-categories based on a Richardson number (Fischeret al., 1979) – radial spreading, offshore spreading, sidedeflecting, diffuse offshore spreading and diffuse shoreimpacting plumes.[5] A detailed classification of the coastal boundary layer

following [Boyce, 1974; Rao and Schwab, 2007] divides thecoastal zone into: (1) a nearshore region that includes the surfand swash zones (2) a frictional boundary layer (FBL) wherebottom and lateral friction effects dominate (3) an inertialboundary layer (IBL) where large-scale oscillations due toinertial effects adjust to the presence of the shoreline and (4)the offshore region that represents the open lake environment.Together the IBL and the FBL are referred to as the coastalboundary layer (CBL) [Csanady, 1972] and represent theregion where flow adjusts to the lateral boundary and transi-tions from the inertial- to the frictional-scale features. Usingcurrent measurements [Murthy and Dunbar, 1981] found thatthe width of the frictional boundary layer (defined as thedistance from the shore to the point of peak kinetic energy)is about 2 km and width of the inertial boundary layer (definedas distance to the point where inertial scales begin to dominateflow) is about 9 km in Lake Huron.[6] Several studies have investigated the dynamics of river

plumes. [Jones et al., 2007] examined mixing of buoyantsurface plumes with ambient waters in idealized flows.Using a combination of dye-studies, ADCP measurements,and surface drifters, mixing due to turbulent diffusion wascalculated for a number of different sites. A detailed reviewof methods used to calculate turbulence in the nearshore isprovided by [Burchard et al., 2008]. [Spydell et al., 2009;Brown et al., 2009], and [Alosairi et al., 2011] have examinedthe effect of shear due to wave-generated alongshore and ripcurrents, on dispersion in the nearshore. Breaking wavesin the surf zone play an important role in wave-dominatedenvironments as quantified by [Pearson et al., 2009]; how-ever, relatively few studies attempted to quantify the effectsof waves on nearshore solute transport, particularly in thecontext of the Great Lakes. Earlier studies [e.g., Bowen andInman, 1974] suggested that mixing across the surf zone ismuch larger than in the alongshore direction and proportionalto (H2/T) whereH is the wave height and T is the wave period.

In relatively low wave-energy environments such as the GreatLakes (e.g., the mean wave height in Southern Lake Michiganduring summer 2008 was 0.2m), the effects of waves andwave-current interactions on solute transport may not beas important as in a marine environment, the focus of severalprevious studies [Svendsen and Putrevu, 1994; Bowen andInman, 1974; Longuet-Higgins, 1970].[7] Shear-augmented diffusion is an important mechanism

that can explain enhanced mixing rates in the nearshore andwas first observed in the Great Lakes by [Csanady, 1966].Due to the combined effects of vertical or lateral shear andturbulence, a dye patch in the nearshore region can spreadat a much faster rate compared to a situation in which mixingis only due to eddy-diffusion driven by turbulence. Inshallow, nearshore areas of the Great Lakes where circulationis primarily wind-driven, the velocity gradients responsible forproducing shear are larger in the vertical compared to those inthe horizontal plane. This shear-augmented diffusion wasinvestigated by [Ojo et al., 2006a, 2006b] and quantified forthe Corpus Christi Bay, Texas using ADCP measurements.They found that shear-augmented diffusion can dominateturbulent mixing in shallow estuarine waters producingeffective mixing rates that are 10 to 20 times higher thanestimates based on turbulence alone. [Chen et al., 2009] iden-tified the interfacial stress term as an important factor affectinglateral spreading of the plume in the near-field. [Hetland,2005] identified that vertical mixing is greatest in the near-field and wind-driven mixing is most important just beyondthe near-field.[8] The horizontal turbulent diffusion coefficient (K) for

oceans and lakes is related to the characteristic length scaleof the flow. [Okubo, 1971] and [Murthy, 1976] used oceano-graphic data to arrive at a power-law relation (1) which esti-mates net turbulent diffusion in the open ocean, away fromboundary effects.

K ¼ aΔb: (1)

[9] Here, the coefficients a and b include the effects ofseveral factors such as wind speed, direction and fetch, surfaceheat fluxes, vertical stratification, current and wave fields, andΔ is the characteristic length-scale. Assuming a homogeneous,isotropic turbulent flow results in the value of b=4/3. The so-called “4/3 power-law” has a theoretical basis followingthe Kolmogorov hypothesis and the work of Batchelor[Batchelor, 1950]. However, homogeneous, isotropic turbu-lence is an idealization rarely found in nature and data reportedfrom field studies produce a value less than 4/3 for b. Forexample, Borthwick [Borthwick, 1980] obtained b = 1.12 inthe surface layer of the Swansea Bay. Equation (1) was foundto describe diffusion (away from any shoreline) in small lakesas well for length scales ranging from 10m to greater than 100m. For example, Lawrence et al. [Lawrence et al., 1995] wereable to fit their tracer data for a small lake in Vancouver,British Columbia using the coefficients a = 3.2 � 10–4 andb = 1.1. In their study to simulate contaminant transport atmultiple scales in the Scheldt River and estuary [de Brauwereet al., 2011] have used a scale dependent diffusion coefficientof K = 0.03Δ1.15.[10] The homogenous and isotropic turbulent field assump-

tions, made in the Okubo relation (1) are a reasonable descrip-tion of mixing in open waters. However, the transport of

THUPAKI ET AL.: DISPERSION IN THE COASTAL BOUNDARY LAYER

1607

tracers in the CBL is affected by the presence of the lateral(shoreline), top (free surface) and bottom (lake bed) boundaries.While large-scale features dominate lake-wide circulation,small-scale features dominate circulation close to the shore.Correctly describing the mixing processes within the CBL isimportant since many transport processes of relevance tocoastal communities primarily occur in this region. [Ojoet al., 2006b] have shown that the presence of shear structureis responsible for significantly enhancing mixing of a tracer inshallow estuarine conditions. When diffusion is controlled byshear, the exponent b in equation (1) is known to approach thelimiting value of 1.0, while the upper limit of 4/3 representsdiffusion in the inertial sub-range [Murthy, 1976].[11] Describing mixing processes in a numerical model of

the CBL can be a challenge due to the range of scalesinvolved. While circulation is primarily wind-driven, verticalmixing across the thermocline is affected by thermal stratifica-tion, and internal waves. Cross-shelf thermal variability due toPoincare waves and a strongly tilted thermocline reported by[Troy et al. [2012]; Wong et al. [2012] also affect nearshorecirculation and transport. From the point of FIB, a transportmodel should have the ability to describe two key aspects ofthe river plumes (both related to hydrodynamics) accurately:(a) advective transport including flow reversals and thefrequency of flow reversals and (b) the nature and strengthof mixing near the shoreline (which control the size and shapeof the river plume). We examine both aspects in the presentpaper using field observations and 3D hydrodynamic andsolute transport modeling.[12] Large-scale circulation in LakeMichigan has been suc-

cessfully described using eddy-resolving hydrodynamicmodels with 2 km grid resolution [Beletsky and Schwab,2001, 2008; Beletsky et al., 1999; Beletsky et al., 2006].Several studies have examined the problem of describingmulti-scale mixing in the nearshore [de Brauwere et al.,2011; Hetland, 2005; Li and Hodgins, 2004; Nekouee,2010]. Vertical mixing is popularly implemented using thek� e or the Mellor-Yamada (MY) 2.5 level [Mellor andYamada, 1982] turbulence closure model. A review ofalgebraic closure models has been provided in [Burchard andBolding, 2001; Burchard et al., 2008]. Horizontal mixing isimplemented using the simpler Smagorinsky model, which hasthe advantage of being computationally stable, robust, and sim-ple to implement. However, the model introduces significantdamping [Moin and Kim, 1982] and improvements such as thedynamic Smagorinsky model and shear-improved Smagorinskymodel have been used in the past [Leveque et al., 2007].[13] In this study, we focus on mixing of a river plume in a

low-wave energy lake environment characterized by flowreversals with a circulation that is primarily wind-driven.Riverine outfalls usually have beaches on either side (i.e., bothupstream and downstream sides relative to the prevailingcurrent direction). Which beach sites are impacted by contami-nated water at any given instant of time depends on the direc-tion of plume travel; therefore flow reversals are an extremelyimportant aspect of FIB transport. One of the objectives of thepaper is to assess the ability of a nested grid nearshore modelbased on the well-documented Smagorinsky and MY formula-tions to accurately describe flow reversals in the nearshore. Weuse velocity measurements from several ADCPs and data froma dye release experiment to evaluate the ability of these formula-tions to describe mixing of a conservative tracer within the CBL.

2. Materials and Methods

2.1. Site Description

[14] The Ogden Dunes beaches are located near Portage,Indiana in Southern Lake Michigan (Figure 1) and frequentlyexperience degraded water quality due to contamination fromthe nearby outfall of Burn Ditch (USGS Gage # 04095090).Field experiments were conducted between May 2008 andSeptember 2008 to study mixing characteristics and solutetransport in the CBL near the outfall. The data from these fieldexperiments form the basis for evaluating the numericalmodels presented in this paper.[15] Five bottom-mounted Acoustic Doppler Current Pro-

filers (ADCPs) were deployed in an upward looking configu-ration. Figure 1 shows the locations of the ADCPs and otherimportant sites in the study area and details of all deploymentsare given in Table 1. The RDI-Monitor and RDI-BBADCPwere deployed in the IBL, while the RDI-Sentinel wasdeployed in the FBL. Additional data were collected in theFBL for a one-week period at locations N1 and N2using two Nortek Aquadopp current profilers. All deployment

Figure 1. (a) Map of southern Lake Michigan showingimportant sites in the field study conducted during summer2008 (b) Finite-difference mesh (2 km resolution) of LakeMichigan used to compute lake-wide circulation. Mesh forthe nearshore model (100 m resolution) is shown in red. (c)Part of the nested nearshore grid used to compute nearshorecirculation and transport showing the bathymetry. The dyerelease location is marked with a red x in the channel in (a).

THUPAKI ET AL.: DISPERSION IN THE COASTAL BOUNDARY LAYER

1608

locations were chosen based on earlier numerical studies[Liu et al., 2006; Thupaki et al., 2010]. Instruments wereprogrammed so that hydrodynamic measurements had a stan-dard deviation of around 0.1 cm/sec [Teledyne, 2006].[16] A continuous-release dye study was conducted on

June 24, 2008 using Rhodamine WT as a tracer, whichwas released into the Burns Ditch outfall (shown in Figure 1)at a constant rate. Concentrations of the dye entering the lakewere measured using a Turner Designs Self-ContainedUnderwater Fluorescence Apparatus (SCUFA) unit mooredat the mouth of the outfall. Plume evolution was trackedby taking multiple transects using a towed SCUFA unitand a hand-held GPS with sub-meter accuracy (LeicaGS20 Professional Data Mapper), on a small motorboat.The procedure was repeated to provide snapshots ofthe plume at two different instants of time separated byapproximately 3 hours in time. Tracer breakthrough data atthe beaches were obtained by taking water samples inknee-deep water, every hour, close to the outfall. Verticaltemperature profile data were collected using NexsensMicro-T temperature loggers attached, at 1 m interval, to asteel cable with a buoy at one end to keep it vertical. Thecable was anchored close to location S.

2.2. Numerical Modeling

[17] Circulation in Lake Michigan was modeled using thehydrodynamic equations in their primitive form derived fromthe Reynolds-averaged form of the Navier-Stokes equationsfor mass and momentum transport. More details about theequations are available in [Vallis, 2006]. The equations weresolved using a modified version of the Princeton OceanModel[Blumberg and Mellor, 1987;Mellor, 1998] that was success-fully used to model large-scale circulation in Lake Michiganby [Beletsky et al., 1999; Beletsky and Schwab, 2001; Beletskyet al., 2006]. The numerical model was adapted to be executedin a nested-grid configuration, so as to resolve small-scalefeatures. Tracer transport was modeled using the unsteady,three-dimensional advection diffusion equation (2).

@C

@tþ u

@C

@xþ v

@C

@yþ w

@C

@z¼ @

@xAH

@C

@x

� �þ @

@yAH

@C

@y

� �

þ @

@zKV

@C

@z

� �(2)

[18] Here AH and KV, which are the eddy diffusivity coeffi-cients in the horizontal and vertical directions, are related tothe eddy viscosity values AM and KM (eddy viscosity in thehorizontal and vertical directions) in the momentum equationsvia the turbulent Prandtl number Pr = AM / AH. Using theSmagorinsky formulation [Smagorinsky, 1963; Pope, 2000]for eddy viscosity in the horizontal directions AM is defined as:

AM ¼ aΔxΔy1

2rV

!þ rV!� �T

��������: (3)

[19] In equation (3), mixing in the horizontal directions (x, y)is assumed to be equal, i.e. anisotropy is neglected. How-ever, computed eddy-viscosity depends on the scale of theunresolved sub-grid scale processes (Δx, Δy). The non-dimensional number (a) has a value around 0.1. Mixing inthe vertical is implemented using the Mellor-Yamada turbu-lence closure model [Mellor and Yamada, 1982]. Thehorizontal directions are discretized using an Arakawa-Cgrid and the vertical is discretized using s-levels that followthe contours of the bathymetry. The momentum equationsare solved using a leapfrog method for the internal modeand the second-order accurate Smolarkiewicz scheme isused to solve for advection. Wind stress at the water surfaceis the prime driver of circulation in Lake Michigan. This iscalculated based on wind speed and direction recorded atmeteorological stations located around Lake Michigan [Liuand Schwab, 1987]. Quality controlled wind measurementsare made available by the National Climatic Data Center(NCDC) and the National Data Buoy Center (NDBC) on theirwebsites. The observed wind datasets were interpolated to thecomputational grid using a nearest neighbor method [Schwaband Morton, 1984]. Seasonal and diurnal changes in tempera-ture of the water column depend on heat flux due to solarinsolation and ambient air temperature. These fluxes arecalculated [McCormick and Meadows, 1988] based on meteo-rological observations recorded at weather monitoring stationsaround Lake Michigan.[20] Circulation near the shore is characterized by small-

scale features, but it is also influenced by large-scale, lake-wide circulation. Large-scale circulation in the hydrodynamicmodel was resolved using a lake-wide grid with a uniformhorizontal resolution of 2 km and 20 s-levels in the verticaldirection. The small-scale features near the shore wereresolved using a nested hydrodynamic model with a horizontalresolution of 100 m and 20 s-levels in the vertical. Resultsfrom the lake-wide hydrodynamic model were interpolatedto provide boundary conditions for the nested model.The lake-wide and nearshore computational grids usedby the hydrodynamic and transport models are shown inFigures 1b and 1c. Energy and mass exchange between thelarge-scale and nearshore models were described as being“one-way interaction”, by ignoring the contributions of near-shore hydrodynamics to lake-wide circulation. Open bound-aries of the nested grid domain were modeled using anupstream advection boundary condition for tracers (tempera-ture, salinity, dye) and a radiation boundary condition forvelocity variables. Surface elevation (hydrostatic pressure)was assigned at the boundary based on interpolated valuesfrom the lake-wide model and the radiation boundary conditionis used to ensure reduced spurious reflections at the boundary.

Table 1. Details of ADCP Deployments and Sampling Locations

Instrument Location CoordinatesPing rate (Ensemble

interval)Depth(m)

600 kHz TRDI-Monitor

M 41.71059 N,87.20996 W

1Hz (5 min) 18.3

600 kHz RDI-BBADCP

B 41.69717 N,87.10078 W

0.1Hz (15 min) 17.6

1200 kHz TRDISentinel

S 41.63813 N,87.18539 W

1Hz (10 min) 9.6

2000 kHz NortekAquadopp

N1 41.66677 N,87.06297 W

1Hz (12 min) 4.1

2000 kHz NortekAquadopp

N2 41.63315 N,87.18839 W

1Hz (12 min) 5.2

Ogden Dunes 1 OD1 41.6279 N,87.1966 W

- -

Ogden Dunes 2 OD2 41.6299 N,87.1875 W

- -

Ogden Dunes 3 OD3 41.6298 N,87.1831 W

- -

THUPAKI ET AL.: DISPERSION IN THE COASTAL BOUNDARY LAYER

1609

3. Results

[21] A comparison of vertically-averaged currents based onobservations and the nearshore hydrodynamic model is shownin Figure 2 for different locations within the CBL. Although thelake-wide model with a 2 km grid resolution (Figure 1b) wasable to resolve the large-scale features of circulation (compari-son not shown), the nested grid model with a 100 m gridresolution provided a better description of the currents closeto the shore. Bottom friction, wind stress at the top surfaceand, internal waves produce a dynamically changing verticalvelocity structure at the location M which is in the IBL. Thevertical variability in alongshore and cross-shore velocity isless prominent at location N2 which is in the FBL. Contourplots of observed and simulated alongshore and cross-shorevelocity profiles at locations M and N2 are shown in Figure 3.A summary of model performance metrics in the form ofRMSE values based on the vertically-averaged currents hasbeen shown in Table 2. The vertical-average values of along-shore and cross-shore currents were calculated by interpolatingsimulation results to observational z-levels and computing themean. The alongshore and cross-shore components of thevelocity were calculated by approximating the coastline to beat an angle of 30o to the East-West direction. Velocity observa-tions from the surface bins were removed and a high-frequencyfilter was used to reduce noise in observed velocities.

[22] Comparison between simulated and observed temper-ature at location S is presented in Figure 4 in which the boxplots denote the variability in temperature within thewater column. The measurement uncertainty for temperature(+/- 1�C) is not shown. Advective transport in the CBL insouthern Lake Michigan is characterized by a strong along-shore current that changes direction frequently. The abilityof the numerical model to predict the reversals in flow direc-tion is shown in Figure 5 by comparing the observed andmodel-predicted times of flow-reversal. Clearly, the nested-grid model with a 100 m resolution was able to capture flowreversals (both timing and their frequency) accurately. Theslight mismatch in the observed and simulated flow-reversaltimes at the location B is attributed to measurement error(the BBADCP was an older ADCP model that was used tomake measurements at this location).[23] Model performance at different scales of motion was

analyzed by comparing the power spectral density (PSD)based on observations and model results. Figure 6 showsthe PSD computed using Welch’s method [Press et al.,2007] for observations and simulation at the locations M, Sand N1. The energy peak at inertial frequency (approxi-mately 18 hr for the latitude of Lake Michigan) was accu-rately predicted by the model. Observations seem to showa second peak around 0.5 cycles per hour (CPH) in theenergy spectrum which was not simulated by the numerical

Figure 2. Comparison of observed (red lines) and simulated (black lines) depth-averaged velocities atdifferent locations within the coastal boundary layer.

THUPAKI ET AL.: DISPERSION IN THE COASTAL BOUNDARY LAYER

1610

model. Least-squares power-law fits for the observed andsimulated velocity power-spectra between 10 CPH and 0.1CPH are also shown in Figure 6. Slopes of the power spectraare over-predicted by the numerical model and energycontained in the smaller scales (higher frequencies) wasnot accurately described by the hydrodynamic model, atboth offshore and nearshore locations. This supports our ear-lier observation that the hydrodynamic model did not repro-duce small-scale features in the water column as accurately.This is similar to the experience of others who have used theSmagorinsky-MY2.5 turbulence closure schemes to simulatenearshore hydrodynamics [Nekouee, 2010].

[24] Circulation in the CBL is anisotropic as a result ofinteractions between large-scale circulation and the shore-line. Flow is predominantly in the alongshore direction as

Figure 3. Comparison of observed and simulated velocity profiles for the alongshore (u) and cross-shore(v) velocities at two different locations in the coastal boundary layer (a) Location M (b) Location N2.

Table 2. Summary of RMSEValues for Current Speeds at DifferentLocations in the CBL

Location RMSE (m/s)

M 0.031S 0.037N1 0.048N2 0.042

Figure 4. Comparison of observed and simulated variabilityin temperatures at location S. The box plots show thevariability in the vertical direction. Uncertainty in themeasurement was �1�C (not shown in the figure).

THUPAKI ET AL.: DISPERSION IN THE COASTAL BOUNDARY LAYER

1611

shown in Figures 2 and 3. Figure 7 shows the anisotropy ofcurrents in the flow by presenting the ratio of two characteristicvelocities u�

v�� �

in the alongshore and cross-shore directionsrespectively. The characteristic velocity u* in the alongshore

direction is defined as u� ¼ffiffiffiffiffiffiffiu02�p

, u0 ¼ u� �u, �u is the vertical

mean of velocity, �u ¼ 1

h

Z h

0u dz where h denotes the water

depth. A similar velocity v*is defined in the cross-shore direc-tion as well. The mean (time-averaged) values of the character-istic velocities in the alongshore and cross-shore directionsare presented in Table 3. Although time-averaged values ofsimulated and observed characteristic velocities are similar,significant differences are noted in their ratios (anisotropy)shown in Figure 7. The mismatch in anisotropy between obser-vations and the simulation can be expected to produce impor-tant differences in comparisons of plume shapes and solutebreakthrough (time series) data.[25] Data from the Rhodamine (RhWT) dye release experi-

ment conducted at the Burns Ditch outfall are compared withresults from the dye-transport model in Figure 8. Due toseveral complicating factors including boat traffic, loss of fluo-rescence due to sunlight, lateral and longitudinal dispersionand retention of tracer mass within the river dead zones, therewas uncertainty in the actual dye concentrations that enteredthe lake at the mouth of the outfall. Assuming an uncertaintyof (�1s) in the values of the RhWT concentrations at themouth of the outfall, uncertainty in the simulated concentra-tions at the Ogden Dunes beaches are included in the compar-isons presented in Figure 8 in the form of error bars. It iscommon practice to calibrate dispersion coefficients in themodel using field data. The dispersion values were adjustedto fit observations by varying the horizontal eddy diffusioncoefficient AH in equation (2). A mean horizontal dispersioncoefficient of 5.6 m2/s was used to simulate the observedRhWT breakthrough at Ogden Dunes beaches, shown inFigure 8. The sampling station OD3, being closest to theoutfall, was most accurately predicted by the model. Movingfurther from the outfall reduces the accuracy of the numerical

model. Because of the complex processes that are responsiblefor mixing and transport of contaminants in the nearshore, it isuseful to examine the spatial extent of the plume as well as thetime series at the individual beaches. Although lack of suffi-cient sampling points prevented a more detailed quantitativecomparison of the spatial extent of the plume, the comparisonsbetween model results and synoptic measurements of theplume shown in Figure 9 indicate that a reasonable agreementwas obtained close to the outfall. Due to the uncertainty ofRhWT concentrations entering the lake, spatial comparisonsof the dye plumes were shown for minimum, average andmaximum loadings entering the lake. The difference betweenobserved and simulated concentrations was greater as onemoved away from the outfall. However, differences werewithin the uncertainty of plume concentrations at the outfallas shown by the time series comparisons presented in Figure 8.

155 160 165 170 175 180 185 190 195 200155

160

165

170

175

180

185

190

195

200

Time of flow reversal (Observed, Julian Day)

Tim

e o

f fl

ow

rev

ersa

l (S

imu

late

d, J

ulia

n D

ay)

BMN1N2S1:1 line

Figure 5. Comparison of observed and simulated times offlow reversals (expressed in Julian days of the year 2008) inthe alongshore direction.

Figure 6. Comparison between observed (red lines) andsimulated (black lines) power spectral densities (PSDs) atdifferent locations within the coastal boundary layer. Thedashed lines represent curve-fits to the data in the frequencyrange 10 CPH to 0.1 CPH. (a, b) Location M (c,d) Location S(e,f) Location N1.

THUPAKI ET AL.: DISPERSION IN THE COASTAL BOUNDARY LAYER

1612

4. Discussion

[26] Our results show that alongshore and cross-shorevelocities close to the shoreline could be resolved using a100 m nested grid with open boundary conditions computedfrom the results of the lake-wide model. Velocity compari-sons with observations, presented in Figures 2 and 3 show thatthe numerical model is able to simulate the hydrodynamicsbetter in deeper waters. As shown in Table 2, model accuracyat observation points located in the IBL (locations M, B) showa significantly lower error relative to comparisons for pointswithin the FBL (locations S, N1, N2). However, flow reversalsdue to changes in wind-direction and large-scale circulation,are accurately simulated by the numerical model (Figure 5)at all locations and this result has important implications formodeling FIB transport in the coastal boundary layer.[27] The power spectral density plots based on observations

and hydrodynamic models show the energy present at aparticular frequency range. This is useful when assessing theperformance of the numerical model in simulating differenttime-scales of motion. The coastal boundary layer is uniquein the sense that lake-wide circulation adjusts to the presenceof lateral boundary features such as riverine outfalls andnatural and artificial features along the coastline. The typicallength-scale of flow features is smaller close the shorelineand energy is transferred from the low-frequency inertialscales to the high-frequency viscous-scales as we move closerto the shoreline. Although the energy peak at the inertialfrequency (approximately 18 hr) is captured well (Figure 6),a second peak around 0.5 CPH is not adequately simulatedby the model. The second peak around 0.5 CPH may be dueto lake-wide free-oscillations in transverse and longitudinaldirections that have been observed in power-spectral analyses

of water level measurements by [As-Salek and Schwab, 2004;Mortimer and Fee, 1976;Rao and Schwab, 1976]. The inabilityof the models to simulate this feature could be due to theinadequate representation of internal and surface waves by thehydrostatic equations solved by the numerical model.[28] As proposed by Kolmogorov’s hypothesis on the

turbulent cascade of energy [Kolmogorov, 1941], the powerspectra can be approximated by a power law relationshipthat is a result of the local equilibrium (at a particular scaleof motion) between the rate of energy transfer and energydissipation. Figure 6 shows the slopes of the power spectracalculated using a least-squares method. The slopes areover-predicted by the numerical model at locations in theIBL as well as FBL. While the energy spectrum is capturedwell in the IBL (Figures 6a, and 6b), the higher dissipationrate in the numerical model results in reduced model accu-racy as one moves into the FBL (Figures 6e, and 6f). As aresult, accuracy of the circulation and transport models alsoreduces from the IBL to the FBL.[29] The nested-grid nearshore numerical model explicitly

resolved the larger scales of motion using a grid size of 100m while the unresolved smaller scales are approximated usingturbulence closure schemes such as the Smagorinsky model.Figure 6 shows that while the model is able to simulate thelarger scales (with time scales of 100 hr to 10 hr), smaller scales(with timescales less than 5 hr) are not accurately described.

Table 3. Observed (Simulated) Characteristic Velocities in theAlongshore and Cross-Shore Directions

Location M S N1 N2

Alongshore characteristicvelocity, u* (m/s)

0.03(0.05) 0.02(0.03) 0.02(0.02) 0.03(0.01)

Cross-shore characteristicvelocity, v* (m/s)

0.03(0.05) 0.02(0.02) 0.02(0.01) 0.02(0.01)

Figure 8. Comparisons between simulated (lines) andobserved (symbols) concentrations of Rhodamine WT at theOgdenDunes beach locations OD1,OD2 andOD3 respectively.

Figure 7. Observed (red) and simulated (black) ratios ofcharacteristic velocities in the alongshore (u *) and cross-shore (v *) directions as a measure of anisotropy at (a) loca-tion M, (b) location S, (c) location N1, and (d) location N2.

THUPAKI ET AL.: DISPERSION IN THE COASTAL BOUNDARY LAYER

1613

[30] Wind stress at the water surface, thermal stratification,and internal waves all result in significant variation in thevelocity profile. The comparisons between vertically-averagedvelocities shown in Figure 2 and the RMSE values in Table 2reveal that, the numerical model is more accurate in theIBL, and the error increases closer to the shoreline. However,the comparison between simulated and observed velocitystructure in Figure 3 shows a distinct top and bottom regionin the IBL, where the velocities are in opposing directions,indicative of internal waves. These features are not presentin the comparisons within the FBL (Figure 3b). The directionsare generally simulated accurately, while the magnitude issometimes over-predicted, especially in the top layer.[31] The hydrodynamic model was able to reproduce the

thermal stratification; however simulated thermocline wastoo diffused similar to the observations made by [Beletskyet al., 2006]. The velocity structure shown in Figure 3 indicatesthat the inertial oscillations observed in the IBL dominate the

vertical variability in alongshore and cross-shore velocities.Closer to the shore, where the epilimnion extends to thelake-bottom, the alongshore and cross-shore velocities in thevertical are nearly uniform. The power spectral density plotsalso show that the inertial peak has dissipated in the FBL.Inaccuracies in simulating the circulation close to theshore maybe attributed to the inability of the model toaccurately simulate dissipation of energy contained withinthe inertial scales.[32] The comparisons between observed and simulated

values of characteristic velocity in the alongshore and cross-shore directions, presented in Figure 7, show the presence ofa highly anisotropic mixing regime in the coastal boundarylayer. The numerical model was able to simulate the meanvalues (Table 3) accurately; however, there are importantdifferences in the peak values that can result in significantdifferences when simulating dye transport. Considering theinaccuracy of model results in the FBL, it is likely that

Figure 9. Comparison of simulated and observed Rhodamine WT concentrations (ppb) at the BurnsDitch outfall for two synoptic measurements within the plume. Due to uncertainty in the concentrationsof the dye entering the lake, simulations are shown for minimum, average and maximum loadings.

THUPAKI ET AL.: DISPERSION IN THE COASTAL BOUNDARY LAYER

1614

numerical models with grid-sizes ofO(10m) or smaller may berequired to reproduce the variability seen close to the shore.However, directly relating grid sizes with resolved frequenciescan be simplistic and misleading. While the large-scale modelhas a grid-size 20 times that of the nearshore model, velocitycomparisons at the location M were found to be comparable.Therefore, improving the representation of the Reynoldsstresses in turbulence model and using more advanced closureschemes, are expected to improve model accuracy in describingthe transfer of energy from larger to the smaller scales. Since themodel used a vertical grid spacing of about 0.5m, insufficientgrid resolution in the vertical direction, is not seen as a reasonfor the model’s inability to simulate observed features accu-rately. Reynolds stresses are apparent stresses due to fluctua-tions in the unfiltered velocity field. The main aim of thevarious turbulence closure models is to find a suitable approx-imation of this term, thereby closing the governing equations.The Smagorinsky closure scheme parameterizes the sub-gridscale turbulent mixing processes based on the gradient/rate-of-strain hypothesis without solving for the turbulent kineticenergy or the length and time scales as is done by more ad-vanced turbulence closure schemes. Momentum and densityeffects dominate mixing and entrainment in the near-fieldand wind-driven mixing processes dominate far-field behaviorof the plume [Hetland, 2005; Nekouee, 2010]. In the absenceof stratification, the time-scale for vertical mixing in the near-shore region is much smaller than mixing in the horizontal di-rections. Plume behavior is therefore dominated by horizontalmixing processes and the horizontal turbulence mixingschemes become increasingly important, especially as onemoves into the far-field.[33] Turbulent diffusion and shear-augmented mixing

dominate transport in the nearshore and adequately representingthe effect of vertical shear, either by resolving the shear struc-ture explicitly or by calibrating the mixing coefficients in themodel against observed data is essential. A mean horizontaldispersion coefficient of 5.6 m2/s was needed to representthe mixing in the nearshore region for the conditions of ourdye release experiment. This is due to the effect of unresolvedshear structure as well as other processes such as surface andinternal waves that were not described by the model. The spa-tial comparison between plumes in the nearshore reveals thatthe model underestimated the plume extent close to the shore.

This could be due to wave-induced alongshore velocity (notaccounted for in the current model) or the lack of anisotropyin the turbulent diffusion field in the model. Finally, an aerialimage of the river plume from the Burns Ditch outfall(Figure 10) shows the complex nature of mixing and thehighly irregular shape of the interface between the river andlake waters. This is in contrast with the smooth and regularplume shapes produced by the advection diffusion equationused in POM (equation (2)). Lagrangian (particle-based)approaches are expected to describe plume shapes better; how-ever, since the success of Lagrangian methods depends on theaccuracy of the hydrodynamic fields on which they are based,it is important to improve fundamental descriptions of hydro-dynamic and wave phenomena in the coastal region in order tomore accurately describe FIB fate and transport. Moreadvanced turbulence closure schemes and coupled wave-current hydrodynamic models are expected to improve theaccuracy of FIB transport models in the surf zone.

5. Conclusions

[34] Field observations and results from a nested-gridnearshore numerical model were used to analyze mixingand transport processes within the coastal boundary layerof southern Lake Michigan. Comparisons between modelresults and ADCP measurements of velocity profiles atdifferent locations within the IBL and FBL show that the hy-drodynamic model was able describe the vertically-averagedvelocity time series as well as vertical velocity profiles ade-quately. Comparisons between the simulated and observedpower-spectral densities show that the turbulence closurescheme was able to describe the transfer of turbulent kinetic en-ergy at the inertial scale, but was unable to accurately describethe energy transfer and dissipation at smaller scales. It was alsoseen that errors in simulated velocities increased as we movefrom the IBL to the shallowwaters, dominated by viscous (fric-tion) effects. The quadratic bottom boundary layer descriptionused in the numerical model, in conjunction with the local(wind-driven) and large-scale forcing, was able to describethe general flow structure, in the vertical including the internalwaves. Direct comparisons with observations were not possiblein the surface layer of the water column due to insufficient ver-tical resolution of the ADCP profiles. The numerical model

Figure 10. An aerial image of the river plume from the Burns Ditch outfall showing complex mixingpatters and the sharp interface between river and lake waters (Photo courtesy of Jerry Mobley).

THUPAKI ET AL.: DISPERSION IN THE COASTAL BOUNDARY LAYER

1615

was able to simulate the temperature of the water column inthe nearshore to within observational accuracy.[35] Inertial waves observed in the IBL were found to

dominate vertical variation in alongshore and cross-shorevelocity components. The observed velocity structures showa great deal of variation in the vertical, particularly in theIBL. However, in the FBL where the thermocline extendsto the lake-bed, the alongshore and cross-shore velocitiesshowed little variation in the vertical. The model’s inabilityto simulate dissipation of internal waves accurately mayhave resulted in higher error closer to the shore. Theother important aspect of the hydrodynamic field in theCBL, flow reversals, is accurately described by the model.Mean values of the anisotropy in the velocity fields (charac-terized by the ratio of the characteristic velocities in thealongshore and cross-shore directions) were predicted rea-sonably well, although a comparison of the time seriesshows considerable mismatch between observed and simu-lated values of anisotropy in the flow field. Small-scale, highfrequency fluctuations, typical of the highly turbulent flowfields, were observed in the measured velocity time-seriesbut the model was unable to reproduce these features perhapsdue to the use of the Smagorinsky turbulence closure modelwhich is known to produce excessive damping.[36] In conclusion, we found that a highly variable vertical

shear structure, small-scale processes and flow reversals areall important features of circulation in the coastal boundarylayer that affect mixing and transport of solutes in thenearshore environment. A three-dimensional hydrodynamicmodel was tested using observations from five ADCPsdeployed during the summer months in southern LakeMichigan. The model was able to simulate the mean velocityfield and energy contained in the larger scales of motion, butsignificantly under-estimated the energy contained in thesmaller scales. Comparisons between the observed andsimulated power-spectral densities showed that turbulentenergy transfer from the large, inertial-scale eddies to thesmaller scales was not accurately described by the model.For a transport model with a resolution of 100m, a mean hor-izontal dispersion coefficient of 5.6 m2/s was required to de-scribe data from our dye release experiment. This value isclosely approximated by the scale-dependent dispersion rela-tion (K = 0.03Δ1.15) proposed by [de Brauwere et al., 2011].

[37] Acknowledgments. This research was funded by the NOAACenter of Excellence for Great Lakes and Human Health. We thank DavidJ. Schwab and Dmitry Beletsky (University of Michigan) for their contribu-tions to this research. We thank Chaopeng Shen, Jie Niu and Kashi Telsangfor their assistance with the 2008 field data collection.

ReferencesAlosairi, Y., J. Imberger, and R. A. Falconer (2011), Mixing and flushing inthe Persian Gulf (Arabian Gulf), J. Geophys. Res. Oceans, 116, C03029.

As-Salek, J. A., and D. J. Schwab (2004), High-Frequency Water LevelFluctuations in Lake Michigan, J. Waterw. Port Coastal Ocean Eng.,130(1), 45–53.

Batchelor, G. K. (1950), The application of the similarity theory of turbulenceto atmospheric diffusion, Q. J. R. Meteorolog. Soc., 76(328), 133–146.

Beletsky, D., and D. J. Schwab (2001), Modeling circulation and thermalstructure in Lake Michigan: Annual cycle and interannual variability,J. Geophys. Res Oceans, 106, 19,745–19,771.

Beletsky, D., and D. J. Schwab (2008), Climatological circulation in LakeMichigan, Geophys. Res. Lett., 2008.

Beletsky, D., J. H. Saylor, and D. J. Schwab (1999), Mean Circulation in theGreat Lakes, J. Great Lakes Res., 25(1), 78–93.

Beletsky, D., D. Schwab, andM.McCormick (2006),Modeling The 1998-2003Summer Circulation And Thermal Structure In Lake Michigan, J. Geophys.Res. Oceans, 111, C10010.

Blumberg, A. F., and G. L. Mellor (1987), A description of a three-dimensional coastal ocean circulation model, In: Three-dimensionalcoastal ocean models, N.S. Heaps (Ed.) pp. 1–16, American GeophysicalUnion, Washington, D.C.

Borthwick, I. (1980), Diffusion and dispesion characteristics in Swansea Bay:Some preliminary results, industrialised embayments and their environmentalproblems: A case study of Swansea Bay. Pergamon, 367–382.

Bowen, A. J., and D. L. Inman (1974), Nearshore mixing due to wavesand wave-induced currents, Rapp. P.-V. Reun.-Cons. Int. Explor. Mer,167, 6–12.

Boyce, F. M. (1974), Some aspects of Great Lakes physics of importance tobiological and chemical processes, J. Fish. Board Can., 31(5), 689–730.

Brown, J., J. MacMahan, A. Reniers, and E. Thornton (2009), Surf zonediffusivity on a rip-channeled beach, J. Geophys. Res. Oceans, 114(C11), C11015.

Burchard, H., and K. Bolding (2001), Comparative analysis of four second-moment turbulence closure models for the oceanic mixed layer, J. Phys.Oceanogr., 31(8), 1943–1968.

Burchard, H., et al. (2008), Observational and numerical modelingmethods for quantifying coastal ocean turbulence and mixing, Prog.Oceanogr., 76(4), 399–442.

Chen, F., D. G. MacDonald, and R. D. Hetland (2009), Lateral spreadingof a near-field river plume: Observations and numerical simulations,J. Geophys. Res. Oceans, 114(C7), C07013.

Csanady, G. T. (1966), Accelerated Diffusion in the Skewed Shear Flow ofLake Currents, J. Geophys. Res. Oceans, 71(2), 411–420.

Csanady, G. T. (1972), The coastal boundary layer in Lake Ontario: Part II.The summer-fall regime, J. Phys. Oceanogr., 2, 166–176.

de Brauwere, A., B. de Brye, P. Servais, J. Passerat, and E. Deleersnijder(2011), Modelling Escherichia coli concentrations in the tidal Scheldtriver and estuary, Water Res., 45, 2724–2738.

Dorfman, M., and K. S. Rosselot (2010), Testing the waters 2010: A Guideto Water Quality at Vacation Beaches, Rep., Natural Resources DefenceCouncil.

Fischer, H. B., E. J. List, R. C. Y. Koh, J. Imberger, and N. H. Brookes (1979),Mixing in Inland and Coastal Waters, Academic Press, San Diego.

Ge, Z., R. L. Whitman, M. B. Nevers, and M. S. Phanikumar (2012a), Wave-induced mass transport affects daily Escherichia coli fluctuations innearshore water, Environ. Sci. Technol., 46(4), 2204–2211, doi:10.1021/es203847n.

Ge, Z., R. L.Whitman,M. B. Nevers,M. S. Phanikumar, andM.N. Byappanahalli(2012b), Nearshore hydrodynamics as loading and forcing factors forEscherichia coli contamination at an embayed beach, Limnol. Oceanogr.,57(1), 362–381, doi:10.4319/lo.2012.57.1.0362.

Hetland, R. D. (2005), Relating river plume structure to vertical mixing,J. Phys. Oceanogr., 35(9), 1667–1688.

Hipsey, M. R., J. P. Antenucci, and J. D. Brookes (2008), A generic,process-based model of microbial pollution in aquatic systems, WaterResour. Res., 44(7), W07408.

Jones, G. R., J. D. Nash, R. L. Doneker, and G. H. Jirka (2007), BuoyantSurface Discharges into Water Bodies. I: Flow Classification and PredictionMethodology, J. Hydraul. Eng. ASCE, 133(9), 1010–1020.

Kolmogorov, N. A. (1941), Dissipation of energy under locally isotropicturbulence, Dolk. Akad. Nauk SSSR, 30, 16–18.

Lawrence, G. A., K. I. Ashley, N. Yonemitsu, and J. R. Ellis (1995),Natural dispersion in a small lake, Limnol. Oceanogr., 40(8),1519–1526.

Leveque, E., F. Toschi, L. Shao, and J. P. Bertoglio (2007), Shear-improvedSmagorinsky model for large-eddy simulation of wall-bounded turbulentflows, J. Fluid Mech., 570(-1), 491–502.

Li, S., and D. O. Hodgins (2004), A dynamically coupled outfall plume-circulation model for effluent dispersion in Burrard Inlet, British Columbia,J. Environ. Eng. Sci., 3(5), 433–449.

Liu, L., M. S. Phanikumar, S. L. Molloy, R. L. Whitman, D. A. Shively,M. B. Nevers, D. J. Schwab, and J. B. Rose (2006), Modeling the transportand inactivation of E. coli and enterococci in the near-shore region of LakeMichigan, Environ. Sci. Technol., 40(16), 5022–5028.

Liu, P. C., and D. J. Schwab (1987), A comparison of methods for estimatingu* from given uz and air-sea temperature differences, J. Geophys. Res.,92, 6488–6494.

Longuet-Higgins, M. S. (1970), Long-shore currents generated by obliquelyincident sea waves: Parts 1 and 2, J. Geophys. Res., 7533, 6778–6801.

McCormick, M. J., and G. A. Meadows (1988), An intercomparison offour mixed layer models in a shallow inland sea, J. Geophys. Res.,93, 6774–6788.

Mellor, G. L. (1998), Users Guide For A Three-Dimensional, PrimitiveEquation, Numerical Ocean Model.

THUPAKI ET AL.: DISPERSION IN THE COASTAL BOUNDARY LAYER

1616

Mellor, G. L., and T. Yamada (1982), Development of a TurbulenceClosure Model for Geophysical Fluid Problems, Rev. Geophys. SpacePhys., 20(4), 851–875.

Moin, P., and J. Kim (1982), Numerical investigation of turbulent channelflow, J. Fluid Mech. 118, 341–377.

Mortimer, C. H., and E. J. Fee (1976), Free surface oscillations and tides ofLakes Michigan and Superior. Philos. Trans. R. Soc. London, A(281), 1–61.

Murthy, C. R. (1976), Horizontal Diffusion Characteristics in Lake Ontario,J. Phys. Oceanogr., 6(1), 76–84.

Murthy, C. R., and D. S. Dunbar (1981), Structure of the Flow withinthe Coastal Boundary Layer of the Great Lakes, J. Phys. Oceanogr.,11, 1567–1577.

Nekouee, N. (2010), NOAA Technical Memorandum GLERL-151:Dynamics and Numerical Modeling of River Plumes in Lakes, edited.

Ojo, T. O., J. S. Bonner, and C. Page (2006a), Studies on turbulent diffusionprocesses and evaluation of diffusivity values from hydrodynamic obser-vations in Corpus Christi Bay, Cont. Shelf Res., 26(20), 2629–2644.

Ojo, T. O., J. S. Bonner, and C. Page (2006b), Observations of shear-augmented diffusion processes and evaluation of effective diffusivityfrom current measurements in Corpus Christi Bay, Cont. Shelf Res.,26(6), 788–803.

Okubo, A. (1971), Oceanic diffusion diagram, Deep Sea Res., 18, 789–902.Pearson, J. M., I. Guymer, J. R. West, and L. E. Coates (2009), SoluteMixing in the Surf Zone, J. Waterw. Port Coastal Ocean Eng. ASCE,135(4), 127–134.

Pope, S. B. (2000), Turbulent Flows, pp. 806, Cambridge University Press,New York.

Press, W. H., S. A. Teukolsky, W. T. Vetterling, and B. P. Flannery (2007),Numerical Recipes: The Art of Scientific Computation, 3rd ed., pp. 1256,Cambridge University Press, New York.

Rao, D. B., and D. J. Schwab (1976), Two-dimensional normal modes in ar-bitrary enclosed basins on a rotating earth: Application to lakes Ontarioand Superior, Philos. Trans. R. Soc. London, A(281), 63–96.

Rao, Y. R., and D. J. Schwab (2007), Transport and Mixing Between theCoastal and Offshore Waters in the Great Lakes: a Review, J. Great LakesRes., 33(1), 202–218.

Rabinovici, S. J. M., R. L. Bernknopf, A. M. Wein, D. L. Coursey,and R. L. Whitman (2004), Economic and health risk trade-offs ofswim closures at a Lake Michigan beach, Environ. Sci. Technol., 38(10),2737–2745.

Schwab, D. J., and J. A. Morton (1984), Estimation of overlake wind speedfrom overland wind speed: Comparison of three methods, J. Great LakesRes., 10, 68–72.

Smagorinsky, J. (1963), General circulation experiments with the primitiveequations, Mon. Weather Rev., 91(3), 99–164.

Spydell, M. S., F. Feddersen, and R. T. Guza (2009), Observations of drifterdispersion in the surfzone: The effect of sheared alongshore currents,J. Geophys. Res. Oceans, 114, C07028.

Svendsen, I. A., and U. Putrevu (1994), Near-shore mixing and dispersion,Proc. R. Soc. London: Math. Phys. Sci., 445(1925), 561–576.

Teledyne, R. D. I. (2006), Self-Contained ADCP Applications: WinSC andPlanADCP User’s Guide, Teledyne, R. D. I.

Thupaki, P., M. S. Phanikumar, D. Beletsky, D. J. Schwab, M. B. Nevers,and R. L. Whitman (2010), Budget Analysis of Escherichia coliat a Southern Lake Michigan Beach, Environ. Sci. Technol., 44(3)pp. 1010–1016, doi:10.1021/es0902232a.

Troy, C. D., S. Ahmed, N. Hawley, and A. Goodwell (2012), Cross-shelfthermal variability in southern Lake Michigan during the stratifiedperiods, J. Geophys. Res., 117(C2), C02028.

Vallis, G. K. (2006), Atmospheric and Oceanic Fluid Dynamics, 745 pp.,Cambridge University Press, Cambridge, U.K.

Wong, S. H. C., A. E. Santoro, N. J. Nidzieko, J. L. Hench, and A. B. Boehm(2012), Coupled physical, chemical, and microbiological measure-ments suggest a connection between internal waves and surf zonewater quality in the Southern California Bight, Cont. Shelf Res., 34,64–78.

THUPAKI ET AL.: DISPERSION IN THE COASTAL BOUNDARY LAYER

1617