journal of great lakes research - queen's university · 2020. 9. 15. · oof q31 hydrodynamics...

11
UNCORRECTED PROOF 1 Hydrodynamics Q3 of eastern Lake Ontario and the upper St. Lawrence River 2 Shastri Q1 Paturi a , Leon Boegman a, , Yerubandi R. Rao b, 1 3 a Department of Civil Engineering, Queen's University, Kingston, ON, Canada K7L 3N6 4 b National Water Research Institute, Environment Canada, Burlington, ON, Canada L7R 5 6 abstract article info 7 Article history: 8 Received 10 January 2011 9 Accepted 27 September 2011 10 Available online xxxx 11 12 Communicated by Joseph Makarewicz 13 14 15 16 Keywords: 17 Hydrodynamics 18 Kingston Basin 19 St. Lawrence River 20 Model calibration 21 Spectral analysis 22 Flow reversal 23 Intake protection zones Q4 24 Eastern Lake Ontario and the upper St. Lawrence River provide drinking water for approximately 175,000 25 people. To understand the ow dynamics surrounding the eight drinking water intakes in this region, the 26 hydrodynamics were simulated using the Estuary and Lake Computer Model (ELCOM) for the period April27 October 2006. Model simulated water levels, temperatures, and current velocities were compared with 28 observations. Root-mean-square errors in temperature and current simulation were ~ 2 °C and 29 ~58 cm s 1 , respectively. Normalized Fourier norms ranged from 0.8 to 1.2. These errors are consistent 30 with other applications of Reynolds-averaged models to the Great Lakes. ELCOM thus reasonably captures 31 the dynamics of the ow regimes in the nearshore region. The ow was found to be predominantly wind 32 induced in the southwestern lacustrine portion of the domain, with observed but not modeled weak near- 33 inertial oscillations, and hydraulically driven in the northeastern riverine portion. Diurnal and semi-diurnal 34 forcing inuenced the ow throughout the domain. Flow reversal of the St. Lawrence River near Kingston 35 occurred during strong easterly storm events. The model results were applied to delineate Intake Protections 36 Zones surrounding the municipal drinking water intakes. 37 © 2011 Published by Elsevier B.V. on behalf of International Association for Great Lakes Research. 38 39 40 41 42 Introduction 43 There are eight municipal drinking water intakes in eastern Lake 44 Ontario and the upper St. Lawrence River (Fig. 1). Of the approximate 45 100,000 m 3 day 1 of water drawn from these intakes, 46% is for do- 46 mestic use servicing approximately 175,000 residents. There are 47 also six wastewater treatment plant outfalls, as well as numerous 48 combined sewers, storm sewers, and industrial outfalls located on 49 the Canadian side of this waterway. 50 Under the Clean Water Act (2006), it was mandated that drinking 51 water intake protection zones (IPZs) be delineated surrounding these 52 intakes, whereby land use will be managed within a zone extending 53 over a 2 h water travel time from each intake. To delineate IPZs, an 54 understanding of the hydrodynamics is required throughout the re- 55 gion, which extends from the Kingston Basin in eastern Lake Ontario 56 to Brockville in the upper St. Lawrence River. 57 The Kingston Basin (Fig. 1) separates the main body of the lake 58 from the entrance to the St. Lawrence River and is characterized by 59 numerous underwater ridges, channels, and islands. Observational 60 studies have shown the existence of complicated hydraulic and 61 wind-induced circulation patterns (Tsanis et al., 1991). Tsanis et al. 62 (1991) found that the ow (0.81.8 cm s 1 ) is strongly inuenced 63 by bottom topography and the circulation is affected by the presence 64 of the islands. For example, the ow around Amherst Island is 180° 65 out of phase with the wind, and the currents exhibit high variability 66 due to the barotropic pressure gradients induced by the wind- 67 driven water level setup. The ow in the Kingston Basin is stratied 68 during summer months. The application of a nite volume model 69 (FVCOM) to Lake Ontario (Shore, 2009) revealed that the ow of 70 water from the main body of Lake Ontario into the Kingston Basin oc- 71 curs when the depth exceeds 20 m and that a reverse ow back into 72 the lake can occur in shallower regions. Shore's (2009) application 73 applied monthly mean meteorological data and had relatively coarse 74 representation of the islands in the region (e.g., Amherst, Wolfe, and 75 Frontenac). Tsanis and Murthy (1990) calculated the outow of 76 water from the Kingston Basin to the St. Lawrence River to be 77 8300 m 3 s 1 . The ow distribution was estimated at 55% in the chan- 78 nel south of Wolfe Island and 45% in the north Channel. 79 The ow through this region of St. Lawrence is controlled by the 80 Moses Saunders hydroelectric dam near Cornwall, ON. Potok and 81 Quinn (1979), applying a one-dimensional transient hydraulic 82 model, found the model ow to be sensitive to the Manning's n 83 roughness coefcient and the ow through the dam. Thompson et 84 al. (2008) applied a two-dimensional (depth averaged) hydrodynam- 85 ic model to the upper St. Lawrence River downstream from Lake 86 Ontario and found that uncertainties in Manning's n contributed to 87 model uncertainties in water levels and velocities. A two- 88 dimensional (depth-averaged) Lagrangian model was also applied 89 to simulate the transport of chemical spills in the upper St. Lawrence Journal of Great Lakes Research xxx (2011) xxxxxx Corresponding author. Tel.: + 1 613 533 6717. E-mail addresses: [email protected] (S. Paturi), [email protected] (L. Boegman), [email protected] (Y.R. Rao). 1 Tel.: +1 905 336 4785. JGLR-00386; No. of pages: 11; 4C: 0380-1330/$ see front matter © 2011 Published by Elsevier B.V. on behalf of International Association for Great Lakes Research. doi:10.1016/j.jglr.2011.09.008 Contents lists available at SciVerse ScienceDirect Journal of Great Lakes Research journal homepage: www.elsevier.com/locate/jglr Please cite this article as: Paturi, S., et al., Hydrodynamics of eastern Lake Ontario and the upper St. Lawrence River, J Great Lakes Res (2011), doi:10.1016/j.jglr.2011.09.008

Upload: others

Post on 15-Oct-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Journal of Great Lakes Research - Queen's University · 2020. 9. 15. · OOF Q31 Hydrodynamics of eastern Lake Ontario and the upper St. Lawrence River Q12 Shastri Paturi a, Leon

UNCO

RRECTED P

RO

OF

1 HydrodynamicsQ3 of eastern Lake Ontario and the upper St. Lawrence River

2 ShastriQ1 Paturi a, Leon Boegman a,⁎, Yerubandi R. Rao b,1

3 a Department of Civil Engineering, Queen's University, Kingston, ON, Canada K7L 3N64 b National Water Research Institute, Environment Canada, Burlington, ON, Canada L7R

5

6

a b s t r a c ta r t i c l e i n f o

7 Article history:8 Received 10 January 20119 Accepted 27 September 201110 Available online xxxx1112 Communicated by Joseph Makarewicz13141516 Keywords:17 Hydrodynamics18 Kingston Basin19 St. Lawrence River20 Model calibration21 Spectral analysis22 Flow reversal23 Intake protection zonesQ4

24Eastern Lake Ontario and the upper St. Lawrence River provide drinking water for approximately 175,00025people. To understand the flow dynamics surrounding the eight drinking water intakes in this region, the26hydrodynamics were simulated using the Estuary and Lake Computer Model (ELCOM) for the period April–27October 2006. Model simulated water levels, temperatures, and current velocities were compared with28observations. Root-mean-square errors in temperature and current simulation were ~2 °C and29~5–8 cm s−1, respectively. Normalized Fourier norms ranged from 0.8 to 1.2. These errors are consistent30with other applications of Reynolds-averaged models to the Great Lakes. ELCOM thus reasonably captures31the dynamics of the flow regimes in the nearshore region. The flow was found to be predominantly wind32induced in the southwestern lacustrine portion of the domain, with observed but not modeled weak near-33inertial oscillations, and hydraulically driven in the northeastern riverine portion. Diurnal and semi-diurnal34forcing influenced the flow throughout the domain. Flow reversal of the St. Lawrence River near Kingston35occurred during strong easterly storm events. The model results were applied to delineate Intake Protections36Zones surrounding the municipal drinking water intakes.37© 2011 Published by Elsevier B.V. on behalf of International Association for Great Lakes Research.

3839

40

41

42 Introduction

43 There are eight municipal drinking water intakes in eastern Lake44 Ontario and the upper St. Lawrence River (Fig. 1). Of the approximate45 100,000 m3 day−1 of water drawn from these intakes, 46% is for do-46 mestic use servicing approximately 175,000 residents. There are47 also six wastewater treatment plant outfalls, as well as numerous48 combined sewers, storm sewers, and industrial outfalls located on49 the Canadian side of this waterway.50 Under the Clean Water Act (2006), it was mandated that drinking51 water intake protection zones (IPZs) be delineated surrounding these52 intakes, whereby land use will be managed within a zone extending53 over a 2 h water travel time from each intake. To delineate IPZs, an54 understanding of the hydrodynamics is required throughout the re-55 gion, which extends from the Kingston Basin in eastern Lake Ontario56 to Brockville in the upper St. Lawrence River.57 The Kingston Basin (Fig. 1) separates the main body of the lake58 from the entrance to the St. Lawrence River and is characterized by59 numerous underwater ridges, channels, and islands. Observational60 studies have shown the existence of complicated hydraulic and61 wind-induced circulation patterns (Tsanis et al., 1991). Tsanis et al.62 (1991) found that the flow (0.8–1.8 cm s−1) is strongly influenced

63by bottom topography and the circulation is affected by the presence64of the islands. For example, the flow around Amherst Island is 180°65out of phase with the wind, and the currents exhibit high variability66due to the barotropic pressure gradients induced by the wind-67driven water level setup. The flow in the Kingston Basin is stratified68during summer months. The application of a finite volume model69(FVCOM) to Lake Ontario (Shore, 2009) revealed that the flow of70water from the main body of Lake Ontario into the Kingston Basin oc-71curs when the depth exceeds 20 m and that a reverse flow back into72the lake can occur in shallower regions. Shore's (2009) application73applied monthly mean meteorological data and had relatively coarse74representation of the islands in the region (e.g., Amherst, Wolfe, and75Frontenac). Tsanis and Murthy (1990) calculated the outflow of76water from the Kingston Basin to the St. Lawrence River to be778300 m3 s−1. The flow distribution was estimated at 55% in the chan-78nel south of Wolfe Island and 45% in the north Channel.79The flow through this region of St. Lawrence is controlled by the80Moses Saunders hydroelectric dam near Cornwall, ON. Potok and81Quinn (1979), applying a one-dimensional transient hydraulic82model, found the model flow to be sensitive to the Manning's n83roughness coefficient and the flow through the dam. Thompson et84al. (2008) applied a two-dimensional (depth averaged) hydrodynam-85ic model to the upper St. Lawrence River downstream from Lake86Ontario and found that uncertainties in Manning's n contributed to87model uncertainties in water levels and velocities. A two-88dimensional (depth-averaged) Lagrangian model was also applied89to simulate the transport of chemical spills in the upper St. Lawrence

Journal of Great Lakes Research xxx (2011) xxx–xxx

⁎ Corresponding author. Tel.: +1 613 533 6717.E-mail addresses: [email protected] (S. Paturi),

[email protected] (L. Boegman), [email protected] (Y.R. Rao).1 Tel.: +1 905 336 4785.

JGLR-00386; No. of pages: 11; 4C:

0380-1330/$ – see front matter © 2011 Published by Elsevier B.V. on behalf of International Association for Great Lakes Research.doi:10.1016/j.jglr.2011.09.008

Contents lists available at SciVerse ScienceDirect

Journal of Great Lakes Research

j ourna l homepage: www.e lsev ie r .com/ locate / jg l r

Please cite this article as: Paturi, S., et al., Hydrodynamics of eastern Lake Ontario and the upper St. Lawrence River, J Great Lakes Res (2011),doi:10.1016/j.jglr.2011.09.008

Page 2: Journal of Great Lakes Research - Queen's University · 2020. 9. 15. · OOF Q31 Hydrodynamics of eastern Lake Ontario and the upper St. Lawrence River Q12 Shastri Paturi a, Leon

UNCO

RRECTED P

RO

OF

90 River (Shen et al., 1995). This model was not comprehensively vali-91 dated against field observations.92 The objective of our study was to simulate the dynamics of both93 the Kingston Basin and upper St. Lawrence River and to apply the94 model to delineate IPZs. This was the first application of a transient95 three-dimensional model to the region, which was directly forced96 with observed meteorology, had detailed representation of the97 islands, and was comprehensively validated against detailed field ob-98 servations. We applied the three-dimensional, hydrostatic Estuary99 and Lake Computer Model (ELCOM; Hodges, 2000) which has been100 shown to accurately simulate the dynamics of stratified lakes and res-101 ervoirs (e.g., Laval et al., 2003). For lake applications, ELCOM outper-102 forms FVCOM in the simulation of the vertical thermal structure103 (Boegman and Rao, 2010; Shore, 2009) and also has the capability104 for water quality simulation (León et al., 2005). Recently, Huang et105 al. (2010) performed basin-wide simulations of Lake Ontario using106 several hydrodynamic models. They found that the modeled lake sur-107 face temperatures from POM (Princeton Ocean Model) and CANDIE108 (Canadian Version of Diecast Model) agreed better with observations109 than ELCOM, which showed a cold bias during summer. They rea-110 soned that this could be due to differences in the parameterization111 of vertical mixing. ELCOM provided better results in the mixed layer112 depth compared to the other models in the offshore areas. All the113 models showed substantial error in simulating subsurface currents.114 The model has not been comprehensively verified in high-resolution115 application to nearshore regions with an open boundary to the116 main lake.117 ELCOM has been previously applied to coupled lake–river systems.118 Morillo et al. (2008) investigated the interaction of two rivers flowing119 into Couer d'Alene Lake and found that basin morphology and wind120 speed and direction have influence on the fate and transport of121 inflowing water from rivers to the lake. Vidal et al. (2005) applied122 ELCOM to Sau reservoir, which was formed by damming the Ter123 River, and found that period of the prevailing wind coincided with pe-124 riod of the third vertical mode. Other applications of ELCOM to reser-125 voirs and reservoirs connected to a river can be found in the literature126 (e.g., Botelho and Imberger, 2007). ELCOM has not been applied to a127 system consisting of the headwaters of a river emanating from a128 large lake, where the water level is controlled by a downstream129 hydraulic structure — as in our application.

130Methods

131Model description

132We applied the Estuary and Lake Computer Model (ELCOM;133Hodges et al., 2000). It is a three-dimensional, hydrodynamic,134z-coordinate model used for predicting the velocity, temperature,135and salinity distribution in natural water bodies subjected to external136environmental forcing such as wind stress, surface thermodynamics,137and inflows and outflows. It is designed to facilitate modeling studies138of aquatic systems over seasonal time scales, although the limit of139computational feasibility depends on the size of the lake, the resolu-140tion requirements, and available computational resources. The141model solves the unsteady, viscous Navier–Stokes equations for in-142compressible flow using the hydrostatic assumption for pressure143(Dallimore and Hodges, 2000). Modeled and simulated processes in-144clude baroclinic and barotropic responses, rotational effects, tidal145forcing, wind stresses, surface thermal forcing, inflows, outflows,146and transport of salt, heat, and passive scalars. The hydrodynamic al-147gorithms in the model are based on the Euler–Lagrange method for148advection of momentum with a conjugate-gradient solution for the149free-surface height (Hodges, 2000). ELCOM has an eddy-viscosity/dif-150fusivity closure scheme for horizontal turbulence correlations and151employs the ULTIMATE QUICKEST advection scheme for scalars. Ver-152tical mixing was computed using an explicit turbulent kinetic energy153budget closure scheme that was applied to each individual water col-154umn of the three-dimensional (3D) flow matrix during each model155time step (Hodges et al., 2000). Turbulent kinetic energy was intro-156duced at the surface from surface wind stress and at the lake bed157through a bottom shear drag coefficient parameterization. The extent158of mixing performed in a single model time step was limited by a159mixing time scale, allowing for partial mixing of vertically adjacent160cells (Laval et al., 2003). Momentum was introduced into the water161column by wind stress at the surface and was distributed vertically162by the 3D mixed-layer model.

163Model setup

164The model grid had a uniform horizontal resolution of165300 m×300 m. The vertical grid had 73 layers with varying resolu-166tion, chosen to reproduce the observed vertical temperature stratifi-167cation with reasonable computational effort (see Hall, 2008;168Boegman and Rao, 2010). The top three layers were at 0.1 m spacing169and the next 40 layers (through the epilimnion and metalimnion) at1700.5 m spacing. The lower layer spacing (in the hypolomnion) ranged171from 1 m to 5 m, at the maximum depth of 70 m. The model grid cov-172ered the region from the Sandhurst Shores intake in the southwest to173the Brockville intake in the northeast (Fig. 1). A grid sensitivity study174using a 200 m×200 m horizontal grid showed negligible differences175in the simulated mean flow relative to the results presented herein.176The bottom boundary was modeled as a turbulent boundary layer.177Stress at the free surface due to wind was modeled as a momentum178source applied to the surface wind-mixed layer (Hodges et al.,1792000). The wind drag coefficient was taken as 1.3×10−3.180The bathymetric grid (Fig. 1) was generated using datasets from181two sources. Bathymetric soundings (331,706 in total) were interpo-182lated for the region between Kingston and Cornwall (Aaron Thomp-183son, Environment Canada). Gridded bathymetric data with 3 s184resolution was used for the region between Kingston and Prince Ed-185ward County. The gridded data were obtained from the National Geo-186physical Data Centre (www.ngdc.noaa.gov). The horizontal datum187was in UTM coordinates, Zone 17, NAD83; the vertical datum was188IGLD 85.189ELCOM was run for a period of 190 days, from 13 April 2006 to 19190October 2006 with a time step of 5 min. Simulations were cold started191(i.e., the fluid was at rest). The initial condition for the temperature

Fig. 1. Map showing the nearshore bathymetric grid with 300 m horizontal resolutionalong with locations of the eight municipal drinking water intakes (1 — Brockville, 2 —

Gananoque, 3 — Kingston Central, 4 — Kingston West, 5 — Fairfield, 6 — Bath, 7 — A.L.Dafoe & 8 — Sandhurst Shores), observational moorings Stations (a — 1262, b — 1263,c— 1264, d— 1265) and the wind data for the stations along St. Lawrence River (Kings-ton airport — e). The solid blue lines are the open boundaries. The wind rose diagramfor Station 1263 is shown in Fig. 2. (For interpretation of the references to color inthis figure legend, the reader is referred to the web version of this article.)

2 S. Paturi et al. / Journal of Great Lakes Research xxx (2011) xxx–xxx

Please cite this article as: Paturi, S., et al., Hydrodynamics of eastern Lake Ontario and the upper St. Lawrence River, J Great Lakes Res (2011),doi:10.1016/j.jglr.2011.09.008

Page 3: Journal of Great Lakes Research - Queen's University · 2020. 9. 15. · OOF Q31 Hydrodynamics of eastern Lake Ontario and the upper St. Lawrence River Q12 Shastri Paturi a, Leon

UNCO

RRECTED P

RO

OF

192 field was provided by the thermister mooring at Station 1263 (Fig. 1).193 The surface boundary conditions were specified, uniformly across the194 domain, using measured overlake meteorological parameters (de-195 scribed in the next section) from Station 1263 in the Kingston basin.196 While spatial variability of winds is likely over the region, two-197 dimensional hydraulic models have shown flows to not significantly198 differ when spatially variable winds were applied as opposed to uni-199 form winds (Aaron Thompson, personal communication). This is like-200 ly because the hydrodynamic response was predominantly wind201 driven in the eastern portion of the model domain (near Station202 1263) and hydraulically driven in the northern section (away from203 Station 1263). Moreover, application of spatially variable meteorolo-204 gy was not computationally feasible with the number of surface grid205 points used in this application.206 The open boundary to Lake Ontario was forced with observed207 water levels from the Kingston gauge station (www.meds-sdmm.208 dfo-mpo.gc.ca) and the observed temperature profile data from Sta-209 tion 1263. Flows at the boundary were computed internally within210 ELCOM from the water level and temperature profile data to ensure211 conservation of volume. Sensitivity analysis (Hall, 2008) showed the212 nearshore simulations to be worse when the model was forced with213 output from a lake-wide model due to model errors in water levels214 and temperature at the location of the open boundary. The St. Law-215 rence River outflow was specified using measurements at Cornwall216 (Moses Saunders hydroelectric dam; Len Falkiner, Great Lakes St.217 Lawrence Regulation Office, Environment Canada), and the effect of218 the control structure on water levels was included by specifying the219 outflow water level from the Brockville gauge station. Flows from220 tributaries into the St. Lawrence River were not included, since they221 account for only 2% of the total river flow (Tsanis et al., 1991).

222 Observational data

223 The forcing data for the model included measured shortwave radi-224 ation (W m−2), net longwave radiation (W m−2), surface air temper-225 ature (°C), wind speed (m s−1) and wind direction (degrees226 clockwise from north), relative humidity, atmospheric pressure227 (Pa), and rainfall amount (mm d−1) and were input at 10 min inter-228 vals. These were measured by the meteorological buoy at Station

2291263 mounted 3.3 m above the water surface. The predominant230wind direction is southwesterly (Fig. 2). Four storm events (winds231>20 m s−1) were observed during days 168, 178, 180, and 192.232Water temperature data were recorded using Onset Tidbit tem-233perature loggers at four thermister chain locations (Stations 1262,2341263, 1264, and 1265; Fig. 1, Table 1). At Station 1262, water current235speed and direction were recorded using a two-axis MAV ultrasonic236current meters. Vertical profiles of currents were obtained at Stations2371263 and 1264 using an upward looking RDI Workhorse Acoustic238Doppler Current Profiler (ADCP) with 1 m vertical bins. Hourly239water levels measured at Kingston and Brockville gauges were also240available.

241Quantification of model error

242Time series of simulated water levels, temperature, and currents243were compared to measured data to assess the model performance.244The ability of the model to reproduce the dominant oscillatory baro-245tropic (surface) and baroclinic (internal) processes was determined246by calculating energy spectra. Root Mean Square (RMS) error was247used to quantify the agreement between model simulated and ob-248served water levels, temperature, and currents:

RMS ¼ 1M∑M

i¼1 f mi −f oi� �2� �1=2

; ð1Þ

249250where fim and fi

o are modeled and observed water levels, temperature251and currents for sample case i (out of M sample cases).252To provide a quantitative comparison between modeled and ob-253served currents, normalized Fourier norms were calculated (e.g.,254Huang et al., 2010):

Fn ¼ 1M∑MΔt

t¼Δt→Vm−

→Vo

��� ���2� �1

2

=1M∑MΔt

t¼Δt→Vo

��� ���2� �1

2 ð2Þ

255256where →Vm and →Vo are modeled and observed currents, respectively.257The smaller Fn, the better the model results fit the observations. Fn258can also be thought of as the relative percentage of variance in the

5%

10%

15%

WEST EAST

SOUTH

NORTH

0 − 2

2 − 4

4 − 6

6 − 8

8 − 10

10 − 12

12 − 14

14 − 16

16 − 18

18 − 20

20 − 24

Wind Speed (ms−1)

Fig. 2. WindQ5 rose diagram at Station 1263. The length of each wedge denotes percentage of occurrence from a particular direction and the color indicates the wind speed. (For in-terpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

3S. Paturi et al. / Journal of Great Lakes Research xxx (2011) xxx–xxx

Please cite this article as: Paturi, S., et al., Hydrodynamics of eastern Lake Ontario and the upper St. Lawrence River, J Great Lakes Res (2011),doi:10.1016/j.jglr.2011.09.008

Page 4: Journal of Great Lakes Research - Queen's University · 2020. 9. 15. · OOF Q31 Hydrodynamics of eastern Lake Ontario and the upper St. Lawrence River Q12 Shastri Paturi a, Leon

UNCO

RRECTED P

RO

OF

259 observed currents that is unexplained by the calculated currents. In260 the case of perfect prediction, Fn=0.

261 Results

262 Water levels

263 The modeled water levels at Kingston and Brockville were com-264 pared to the observed gauge water level data (Fig. 3). The modeled265 water level at Brockville (Fig. 3a) is slightly larger in amplitude than266 the observed; the RMS error between the modeled and the observed267 was 4.5 cm. However, at Kingston (Fig. 3b) the modeled water level268 amplitude was consistent with the observed and the RMS error was269 1.1 cm. Modeled results were consistent in phase. The favorable com-270 parisons are not unexpected because the gauge data used for compar-271 ison here, which are in close proximity to the data shown, were also272 applied to force the open boundaries.

273Spectral energy plots of the water levels at Kingston (Fig. 4a) dem-274onstrated statistically significant peaks in the model results and field275data at 24 h, 12 h, and at the periods of the first (5.06 h), second276(3.21 h), and third (2.32 h) longitudinal surface seiche modes277(Hamblin, 1982). However, the modeled spectra showed a peak for278the fourth (1.71 h) longitudinal seiche mode. This peak was not279seen in the observed spectra as the data sample frequency was 1 h280(Hamblin, 1982). At Brockville (Fig. 4b) the significant peaks repre-281sented the 24 h, 12 h and the first and the third longitudinal surface282seiche modes. The 24 h peak was due to the wind forcing (Fig. 4c).283There was also evidence of a broad ~10-day peak in the wind and284water level spectra, resulting from frontal storm systems characteris-285tic to the region (Hamblin, 1987). The model underestimated these286low frequency (5–10 days) surface oscillations caused by wind-287induced set-up of the free-surface, potentially due to the constant288surface drag formulation. At much lower energy levels, significant289peaks corresponding to the first (45 min) transverse seiche mode

Table 1t1:1

Table listing details of deployment at the mooring stations by NWRI, Environment Canada. All the moorings were deployed in 2006. The accuracy of the thermister was ±0.2 °C, andthat of the MAV and the ADCP was ±0.3 cm s−1and ±0.25 cm s−1. The data measured from MAV was averaged from a 60 min burst of 128 2Hz samples.

t1:2t1:3 Station Parameter Instrument Deployment period (times in GMT) Sampling interval (min) Depth of measurement (m)

t1:4 1262 Temperature Onset Tidbit temperature logger 12 April (12:00)–26 July (02:10) 10 [1 3 5 7 9.5 13 15 16.5]t1:5 27 July (13:50)–15 Nov (16:10)t1:6 Current MAV 12 April (12:00)–26 July (14:00) 60 [11]t1:7 12 April (12:00)–26 July (12:00)t1:8 1263 Temperature Onset Tidbit temperature logger 12 April (12:00)–26 July (13:00) 10 [1 3 5 7 10.8 12 14 16]t1:9 27 July (15:20)–19 Oct (12:30)t1:10 Current ADCP 12 April (12:00)–26 July (12:00) 30 [0 0.5:1:15.5]t1:11 Meteorological MET buoy 11 April–26 Dect1:12 1264 Temperature Onset Tidbit temperature logger 12 April (12:00)–26 July (17:10) 10 [1 3 5 7 10 12 14 15]t1:13 27 July (12:00)–15 Nov (18:10)t1:14 Current ADCP 27 July (12:00)–15 Nov (15:00) 30 [0:15]t1:15 1265 Temperature Onset Tidbit temperature logger 12 April (12:00)–26 July (16:00) 10 [1 3 5 7 9.5 13 15 16.5]t1:16 27 July (12:00)–15 Nov (17:40)

120 140 160 180 200 240220 260 280

120 140 160 180 200 240220 260 280

−0.2

−0.1

0

0.1

0.2

0.3

Hou

rly S

urfa

ce W

ater

Lev

el (

m) Brockville Water Level

a

−0.1

0

0.1

0.2

0.3

Time (days)

Hou

rly S

urfa

ce W

ater

Lev

el (

m) Kingston Water Level

bObserved

Modeled

Fig. 3. Comparison of hourly water level data measured at (a) Brockville and (b) Kingston gauges with hourly modeled output of water surface height at Brockville and Kingston,respectively.

4 S. Paturi et al. / Journal of Great Lakes Research xxx (2011) xxx–xxx

Please cite this article as: Paturi, S., et al., Hydrodynamics of eastern Lake Ontario and the upper St. Lawrence River, J Great Lakes Res (2011),doi:10.1016/j.jglr.2011.09.008

Page 5: Journal of Great Lakes Research - Queen's University · 2020. 9. 15. · OOF Q31 Hydrodynamics of eastern Lake Ontario and the upper St. Lawrence River Q12 Shastri Paturi a, Leon

UNCO

RRECTED P

RO

OF

290 (Denison, 1908) are found in the modeled Kingston spectra but the291 second (22.5 min) transverse seiche mode was not present. This292 was likely associated with a north–south barotropic oscillation.

293 Temperatures

294 Comparisons between modeled and observed temperatures north295 of Amherst and Wolfe Islands at Stations 1262 (Fig. 5a,e) and 1264296 (Fig. 5c,g), respectively showed the modeled mixed layer depth to297 be shallower than observed. This resulted in a cold bias during the298 summer of ~2 °C. Overall the model captures the seasonal evolution299 of the temperature profile.300 Stations 1263 (Fig. 5b,f) and 1265 (Fig. 5d,h) showed a sharp ob-301 served thermocline at depth 15 m. The modeled thermocline was302 not as abrupt as observed, potentially due to numerical diffusion303 (Laval et al., 2003) and for the discrete nature of the vertical grid res-304 olution. The warming of the water column, temporal evolution of the305 stratification, erosion of stratification, and fall cooling were also well306 simulated at these locations.307 The RMS error between modeled and observed temperatures was308 typically 1–2 °C (Table 2) with a maximum of 3 °C (Fig. 6a,b,c,d). A309 cold bias of ~2 °C in the model result was seen at Stations 1262 (Fig.310 6a), 1263 (Fig. 6b) and 1264 (Fig. 6c). The maximum error (2–3 °C)311 was observed through the thermocline, where the strong tempera-312 ture gradient and phase differences between observed and modeled313 internal waves caused small errors in model skill to lead to larger314 RMS errors (Dorostkar et al., 2010). These results were consistent315 with the application of ELCOM and other models to temperate lakes316 (e.g., Rao et al., 2009; Huang et al., 2010), which have shown maxi-317 mum errors through the thermocline.318 Vertical mode one baroclinic oscillations (internal waves) were319 evident in the modeled and observed temperature profiles. The320 strength of these motions may be quantified using the integrated

321potential (IPE) energy of the water column (Antenucci and322Imberger, 2001). Fig. 7a,b shows IPE spectra for Stations 1262 and3231263, respectively, where 24 h and 12 h oscillations were evident in324the spectra, with the exception of Station 1263 where the significant325energy peak was at the inertial period (~17 h). The lack of an inertial326peak in the modeled data likely resulted from the model domain327being too small to completely resolve internal Poincare waves, and328these waves were not propagating to the domain through the open329boundary (see Schwab, 1977 and Rao and Schwab, 2007). The IPE330spectral peaks at Stations 1264 and 1265 (Fig. 7c,d) also showed331peaks at 24 h and 12 h. These two stations were at the confluence of332the St. Lawrence River and the lake, and the peaks resulted from diur-333nal processes (e.g., sea-breeze winds and day-to-night heating and334cooling).

335Currents

336A comparison between modeled and observed, east (Fig. 8a) and337north (Fig. 8b) components, MAV currents at Station 1262 shows338that the current direction was consistent with the field observations.339The RMS errors for the east and north component velocities were3405.7 cm s−1 and 3.8 cm s−1, respectively (Table 2). The model did341not appear to be systematically over/under estimating current veloc-342ities, and the RMS errors appeared (Fig. 8) to result from error in343modeled direction as opposed to speed.344Comparison of the modeled east–west current profiles at Station3451263 (Fig. 9a,e) reveals that the directions were well modeled, but346the model occasionally overestimated the depth to which the strong347surface currents penetrated (e.g., day 135 and day 190); comparison348of the observed and modeled north–south current profiles at Station3491263 (Fig. 9b,f) shows similar results (e.g., days 185 and 190). These350differences may be due to the topographic differences between the351300 m bathymetric grid and the actual bathymetry (Hall, 2008). The

Frequency (Hz)

(a) − Kingston

ObservedModeled24hr & 12hrLong. seichesTrans. seiches

Frequency (Hz)

(b) − Brockville

ObservedModeled24hr & 12hrLong. seichesTrans. seiches

10−410−6

10−4

104

102

100

10-2

10-4

10-6

104

102

100

10-2

10-4

10-6

10−6 10−410−6

106

104

102

Frequency (Hz)

Spe

ctra

l Ene

rgy

(m2 H

z−1 )

Spe

ctra

l Ene

rgy

(m2 H

z−1 )

Spe

ctra

l Ene

rgy

(m2 H

z−1 )

(c) − 1263 Wind Speed

24 hr

Fig. 4. (a) Spectra of water level at Kingston (observed and modeled), (b) Spectra of water level at Brockville (observed and modeled) and (c) Spectra of wind speed collected(10 min) at Station 1263 (observed). The red dotted line is the 95% confidence level.Q2

5S. Paturi et al. / Journal of Great Lakes Research xxx (2011) xxx–xxx

Please cite this article as: Paturi, S., et al., Hydrodynamics of eastern Lake Ontario and the upper St. Lawrence River, J Great Lakes Res (2011),doi:10.1016/j.jglr.2011.09.008

Page 6: Journal of Great Lakes Research - Queen's University · 2020. 9. 15. · OOF Q31 Hydrodynamics of eastern Lake Ontario and the upper St. Lawrence River Q12 Shastri Paturi a, Leon

UNCO

RRECTED P

RO

OF

352 RMS error magnitude for the east–west and north–south velocity353 components was 4.0 cm s−1 and 5.1 cm s−1 respectively (Table 2).354 At Station 1264, the hydraulic flow through the St. Lawrence River355 was evident with observed and modeled currents having an easterly356 mean flow component of ~5–10 cm s−1 (Fig. 9c,g). A baroclinic ve-357 locity structure in the modeled data, with a flow reversal about the358 mid-water column (~5 m depth), is not observed. During this time,359 the water column was fully mixed (Fig. 5c), and so these flow dynam-360 ics did not result from an internal wave but were likely due to wind-361 induced and hydraulic flow being in opposite directions. The modeled362 north–south velocity (Fig. 9h) was over-estimated in magnitude at363 mid-depth with a strong southerly component in comparison to the364 observed velocity (Fig. 9d) but was consistent in direction.365 A wind-induced flow reversal of the St. Lawrence River was ob-366 served near days 237–247, 255, and 258 (Fig. 10b), resulting from367 strong (~10 m s−1) easterly winds (Fig. 10a). A combined sewer

368overflow occurred around day 245, making this reversal important369for source water protection, as City of Kingston Municipal Engineers370were not aware of these flow dynamics, which have the potential to

Time (days)

Dep

th (

m)

(a) − Obs 1262

100 150 200 250

0

5

10

15

(e) − Mod 1262

100 150 200 250

0

5

10

15

(b) − Obs 1263

100 150 200 250

0

5

10

15

(f) − Mod 1263

100 150 200 250

0

5

10

15

(c) − Obs 1264

100 150 200 250

0

5

10

15

(g) − Mod 1264

100 150 200 250

0

5

10

15

(d) − Obs 1265

100 150 200 250

0

5

10

15

(h) − Mod 1265

100 150 200 250

0

5

10

15

Tem

pera

ture

(o C

)

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

Fig. 5. 10 min observed (a, b, c, d) and modeled (e, f, g, h) temperature profile comparisons at Stations 1262, 1263, 1264 and 1265. The times are in EDT.

Table 2t2:1

Mean RMS (Root Mean Square) error between model and observation for temperatureand currents and Normalized Fourier norms (Fn) for currents.

t2:2t2:3 Station Temperature

RMS (°C)Current RMS (cm s−1) Fn

t2:4 East North

t2:5 1262 1.86 5.68 3.80 0.8t2:6 1263 1.45 3.98 5.10 1.2t2:7 1264 1.50 6.30 2.61 1.0t2:8 1265 1.45 – – –

RMS Error (oC)4

0

2

4

6

8

10

12

14

16

Station 1265d

4

0

2

4

6

8

10

12

14

16

Station 1264c

4

0

2

4

6

8

10

12

14

16

Station 1263b

0 20 20 20 2 4

0

2

4

6

8

10

12

14

16

Dep

th (

m)

Station 1262a

Fig. 6. RMS error profiles for temperature at (a) Station 1262, (b) Station 1263,(c) Station 1264, and (d) Station 1265.

6 S. Paturi et al. / Journal of Great Lakes Research xxx (2011) xxx–xxx

Please cite this article as: Paturi, S., et al., Hydrodynamics of eastern Lake Ontario and the upper St. Lawrence River, J Great Lakes Res (2011),doi:10.1016/j.jglr.2011.09.008

Page 7: Journal of Great Lakes Research - Queen's University · 2020. 9. 15. · OOF Q31 Hydrodynamics of eastern Lake Ontario and the upper St. Lawrence River Q12 Shastri Paturi a, Leon

UNCO

RRECTED P

RO

OF

371 rapidly transport waste up-river toward drinking water intakes. The372 model reproduced the flow reversals, but the observed currents373 were stronger at depth and the modeled currents were elevated374 near the surface (Fig. 10b,c). These differences may result from spatial375 variability in the surface winds, which were not implemented in the376 surface forcing. Uniform winds have been applied from Station377 1263, but these may be sheltered from the east by Simcoe Island.378 Winds further along the St. Lawrence River are channeled in an379 east–west direction through the river valley (Fig. 1). Overall, at Sta-380 tion 1264 the model underestimated the north–south current magni-381 tude by ~4 cm s−1 in the surface layer (Fig. 9d,h).

382The water column RMS error profiles for current magnitude at Sta-383tions 1263 (Fig. 11a) and 1264 (Fig. 11b) showed that model currents384were much stronger in magnitude than observed, especially at the385surface. The RMS errors were between 1 and 10 cm s−1. The east386component velocity RMS error at Station 1264 was higher near the387bottom due to the riverward flow in the bottom layers around388Wolfe Island (Tsanis et al., 1991). Computed Fourier norms ranged389from 0.8bFnb1.2 (Table 2). The errors were slightly larger in our390nearshore simulations compared to lake-wide simulations (Huang391et al., 2010) where 0.4bFnb0.9. This could be due to the complicated392topography and presence of numerous islands in the present model

Frequency (Hz)

Spe

ctra

l ene

rgy

(m2

Hz−

1 )

a b ObservedModeled24 hrInertial12hr

c

10−8 10−6 10−4 10−210−8 10−6 10−4 10−2

10−8 10−6 10−4 10−210−8 10−6 10−4 10−2

10−5

100

105

1010

1015

10−5

100

105

1010

1015

10−5

100

105

1010

1015

10−5

100

105

1010

1015

d

Fig. 7. Integrated Potential Energy (IPE) of the observed and modeled temperature data at (a) Station 1262, (b) Station 1263, (c) Station 1264, and (d) Station 1265.

120 140 160 180 200 220 240 260−20

−10

0

10

20

Spe

ed (

cms−

1 )

East−velocity

a

120 140 160 180 200 220 240 260−20

−10

0

10

20

Time (days)

Spe

ed (

cms−

1 )

North−velocity

b

Observed

Modeled

Fig. 8. Daily averaged observed and modeled (a) East (U — component) and (b) North (V — component) velocity comparisons at 11 m depth at Station 1262. Times are in EDT.

7S. Paturi et al. / Journal of Great Lakes Research xxx (2011) xxx–xxx

Please cite this article as: Paturi, S., et al., Hydrodynamics of eastern Lake Ontario and the upper St. Lawrence River, J Great Lakes Res (2011),doi:10.1016/j.jglr.2011.09.008

Page 8: Journal of Great Lakes Research - Queen's University · 2020. 9. 15. · OOF Q31 Hydrodynamics of eastern Lake Ontario and the upper St. Lawrence River Q12 Shastri Paturi a, Leon

UNCO

RRECTED P

RO

OF

393 domain (see discussion in Hall, 2008). In Lake Michigan, Beletsky et394 al. (2006) reported 0.55bFnb1.59.

395 Case study — IPZ delineation

396 Given the above-mentioned importance of local hydrodynamics397 on spill transport in the upper St. Lawrence River, we present sample398 IPZ delineations for the Kingston region. Intake Protection Zone One399 (IPZ-1) was defined at a 1 km radius around an intake, and an IPZ-2400 was defined as the distance from each intake where a fluid parcel401 would undergo a 2-hr travel time to reach the intake during 10-yr402 storm conditions (Ontario Ministry of the Environment, 2005). The403 2-h travel time was requested by the drinking water treatment404 plant operators so that they could shut down the intake in the case405 of a spill. To reproduce these conditions, ELCOM was run with surface406 winds scaled such that the maximum gust observed during the spring407 (21.5 m s−1on day 168), summer (24.0 m s−1 on day 192), and fall408 (15.6 m s−1 on day 286) seasons matched 10-yr conditions409 (23.3 m s−1) as observed at the Kingston Airport (Fig. 1) located410 nearby on the shore of Lake Ontario (National Building Code of411 Canada, 2005). Wind events were scaled during each season to ac-412 count for seasonal differences in wind direction, lake level, and strat-413 ification. The above validation of the model hydrodynamics to414 observed data provides faith in the model results, and these results415 are used to delineate the IPZs.

416The 10-yr, 1-hr average wind gusts at Kingston Airport were mea-417sured at 10 m above the surface, and so were scaled from 10 m to the4183.3 m height of the Station 1263 wind anemometer using a power law419relation (Schertzer, 1987), giving 19.89 m s−1. The winds were then420adjusted over 24 h storm duration (maximum gust plus/minus42112 h) such that the maximum 1-hr average wind gust matched the42210-yr condition corrected to 3.3 m. This procedure was repeated for423the spring, summer, and fall storm events. IPZ delineations were424then constructed at each intake using the 2-hr reverse progressive425vector (RPV) diagrams calculated from the modeled surface currents426at the eight locations during the 24 h storm events. RPVs were427deemed sufficient to capture the flow dynamics surrounding each in-428take because the modeled velocities did not change dramatically429within 1–2 km of each intake.430RPV diagrams are similar in principle to reverse particle tracking,431whereby fluid parcels continuously released at the intake are tracked432backward in time, such that their position within the flow domain 2 h433prior to release may be determined. The RPV diagram for Kingston434Central (Fig. 12a) and West (Fig. 12b) intake shows the effect of sea-435sonal variability in wind direction and storm intensity on transport of436fluid parcels near the intake. It can be seen that the storm event dur-437ing fall was much weaker (RPV is ≤2 km) than during spring and438summer at both the intakes, with a change in direction at the Kings-439ton Central intake. The fall storm event showed a more eastwardly440movement of water parcels at the Kingston Central intake. The IPZ-

Time (days)

Dep

th (

m)

(a) − Obs 1263 (E)0

5

10

15

(e) − Mod 1263 (E)0

5

10

15

(b) − Obs 1263 (N)0

5

10

15

(f) − Mod 1263 (N)

100 120 140 160 180 200

100 120 140 160 180 200

100 120 140 160 180 200

100 120 140 160 180 200

0

5

10

15

(c) − Obs 1264 (E)0

5

10

15

(g) − Mod 1264 (E)0

5

10

15

(d) − Obs 1264 (N)0

5

10

15

(h) − Mod 1264 (N)

210 220 230 240 250 260 270 280 290

210 220 230 240 250 260 270 280 290

210 220 230 240 250 260 270 280 290

210 220 230 240 250 260 270 280 290

0

5

10

15

Spe

ed (

cms−

1 )

−4 0

−3 0

−2 0

−1 0

−5

0

5

10

20

30

40

Fig. 9. Daily averaged observed (a — East, b — North) and modeled (e — East, f — North) velocity comparisons at Station 1263, and observed (c — East, d — North) and modeled(g — east, h — north) velocity comparisons at Station 1264. Times are in EDT.

8 S. Paturi et al. / Journal of Great Lakes Research xxx (2011) xxx–xxx

Please cite this article as: Paturi, S., et al., Hydrodynamics of eastern Lake Ontario and the upper St. Lawrence River, J Great Lakes Res (2011),doi:10.1016/j.jglr.2011.09.008

Page 9: Journal of Great Lakes Research - Queen's University · 2020. 9. 15. · OOF Q31 Hydrodynamics of eastern Lake Ontario and the upper St. Lawrence River Q12 Shastri Paturi a, Leon

UNCO

RRECTED P

RO

OF

441 2 was delineated as the maximum extent of the 2-hr RPV diagrams442 surrounding each intake. The error associated with the IPZ-2 can be443 estimated from the model error in simulating current velocity. At444 the surface (Fig. 11a,b) the RMS errors ranged from 4 to 10 cm s−1,445 which corresponds to an error of around 300–700 m in the IPZ-2,446 over the 2-hr travel time.447 Offshore Kingston, the Kingston Central, and West intake IPZ-2448 (Fig. 12c) are oriented parallel to the shore. This is consistent with449 the primary flow from the south-west along the St. Lawrence River450 hydraulic axis, which is also coincident with the predominant wind451 direction (Fig. 2) and the shore parallel boundary condition required452 by inertial currents (Rao and Schwab, 2007). Reversal of the St.

453Lawrence flow is seen during easterly storm events that cause the454IPZs, at both these locations, to overlap each other and include the455majority of the Kingston waterfront, with the exception of Cataraqui456Bay. Similar IPZ-2 was computed for the six other intakes in the457Cataraqui Region Conservation Authority jurisdiction (Paturi and458Boegman, 2008). In the upper St. Lawrence River, close to Gananoque459and Brockville, the IPZ-2 showed a strong polarization from the hy-460draulic riverine flow. At Gananoque, the effect of flow reversal during461easterly wind events was reduced, relative to offshore Kingston, due462to wind sheltering and reduced fetch lengths.

463Discussion and conclusions

464The ELCOM model was applied to simulate the hydrodynamics in465eastern Lake Ontario and the upper St. Lawrence River and the results466were comprehensively validated against observations collected dur-467ing April–October, 2006. Spectral analysis revealed a significant468wind-induced circulation in the Kingston Basin; consistent with the469previous studies (Tsanis et al., 1991; Boegman and Rao, 2010) in the470region. Strong interconnectivity between the lacustrine and riverine471systems is evident. Wind and lake seiche effects are found in the St.472Lawrence River as far as Brockville and flow regulation from the473dam at Cornwall is evident at Wolfe Island. Here a riverward flow of474the bottom water also occurs. The transient hydrodynamics of the re-475gion are thus mostly wind influenced, and storm events result in flow476reversals, which can have significant impacts on the transport of477contaminants.478To accurately model water levels throughout the domain, the in-479fluence of the downstream control structure at Cornwall was480accounted for, by specifying both the flow rate and water level at481the outflow boundary. This suggests that the dam is regulating the482upstream water level causing an M1 flow profile in the river, where483the flow depth is greater than the uniform depth that would occur484without the downstream control structure. Overall, our water level485simulations were 10 times more accurate than the ±0.16 m error486reported using a depth-averaged model (Thompson et al., 2008).

−10

0

10S

peed

(m

s−1 ) Wind (1hr)

East component

North component

Bypass

Observed Data

Dep

th (

m)

0

5

10

15

Modeled Data

Time (days)

Dep

th (

m)

235 240 245 250 255 260

235 240 245 250 255 260

235 240 245 250 255 260

0

5

10

15

Spe

ed (

cms−

1 )

−40

−30

−20

−10

−5

0

5

10

20

30

40

a

b

c

Fig. 10. (a) Hourly averaged east and north components of wind speed collected at Station 1263. The shaded region is the Combined Sewer Outflow event (b) 30 min east velocitycomponent measured at Station 1264 and (c) 30 min modeled east velocity component at Station 1264.

RMS Error (cms−1)

8 100 2 4 6 128 100 2 4 6 12

Station 1264b

EastNorth

0

2

4

6

8

10

12

14

16

Dep

th (

m)

0

2

4

6

8

10

12

14

16

Dep

th (

m)

Station 1263a

EastNorth

Fig. 11. RMS error profile for currents (east and north velocity components) at(a) Station 1263 and (b) Station 1264.

9S. Paturi et al. / Journal of Great Lakes Research xxx (2011) xxx–xxx

Please cite this article as: Paturi, S., et al., Hydrodynamics of eastern Lake Ontario and the upper St. Lawrence River, J Great Lakes Res (2011),doi:10.1016/j.jglr.2011.09.008

Page 10: Journal of Great Lakes Research - Queen's University · 2020. 9. 15. · OOF Q31 Hydrodynamics of eastern Lake Ontario and the upper St. Lawrence River Q12 Shastri Paturi a, Leon

UNCO

RRECTED P

RO

OF

487 Kelvin waves were not observed or modeled within the domain.488 This was expected because lake-wide simulations (Boegman and489 Rao, 2010) have shown that the Kelvin wave does not propagate490 from the main lake-basin into the shallower Kingston Basin over the491 Duck-Galoo Ridge. A weak near-inertial Poincaré wave signal was ob-492 served at Station 1263. This mooring is within the frictional boundary493 layer, and consequently a strong Poincaré wave signal was not494 expected (Rao and Schwab, 2007); the inertial circles became shore495 parallel and strongly damped to satisfy the no-flux boundary condi-496 tion. Moreover, the single mooring used for the open-lake boundary497 condition was insufficient to propagate basin-scale inertial oscilla-498 tions into the model domain (Schwab, 1977).499 Previous modeling of the Kingston Basin flow, with FVCOM, has500 shown a favorable comparison between modeled and observed sum-501 mer averaged currents during 1986–87 (Tsanis et al., 1991; Shore,502 2009). Comparisons between our 2006 results and these data are503 not in direct agreement, differing in magnitude but agreeing in direc-504 tion. These differences may easily result from the variability in wind505 forcing events and river flows for the different simulation years. How-506 ever, some similarities are evident; for example, both ELCOM and507 FVCOM simulated a gyre in the Bay of Quinte region (Fig. 13a).508 Gyres are also simulated by ELCOM in the channel north of Amherst509 Island, north off Wolfe Island, and south of Kingston Basin (between510 Amherst Island and Wolfe Island). Tsanis et al. (1991) observed that511 the 24 h peak in the spectral signal of water currents around Amherst512 Island was coherent with the wind but was 180° out of phase with the513 wind direction. We found a similar 24 h peak in the wind and IPE514 spectra at nearby Station 1264, which could be due to sea-breeze ef-515 fects. Similarly, the 24 h signals at Station 1265 could result from516 daily river flow regulation by the hydroelectric dam at Cornwall517 (Tsanis and Murthy, 1990). The averaged currents at 20 m (Fig.518 13b) showed a return flow from Kingston Basin into Lake Ontario,519 similar in direction but smaller in magnitude, in comparison to520 1986–87 data by Tsanis et al. (1991) and from the application of521 FVCOM by Shore (2009).

522Overall, the modeled hydrodynamics agree well, quantitatively523and qualitatively with field observations. The RMS error in tempera-524ture is approx 1–2 °C through the epilimnion, when the thermocline525deepens during summer. The current speed errors are ~5–8 cm s−1

526and appear, at times, to be a result of inconsistencies in flow direction527as opposed to momentum transfer from the wind. Current velocity528errors are mostly found at exposed regions with large fetch lengths,529Stations 1263 and 1264, during strong wind events and consequently

−6 −4 −2−4

−3

−2

−1

0

1

2

Easting (km)

Nor

thin

g (k

m)

−6 −4 −20 2 0 2−4

−3

−2

−1

0

1

2

Easting (km)

1264

Longitude

Latit

ude

38’ 36’ 76oW 32’ 30’ 28’

11’

12’

44oN

14’

15’

Kingston Central

Kingston West

a b

c

Fig. 12. Reverse Progressive Vector (RPV) diagram for (a) Kingston Central intake and (b) Kingston West intake for spring (red), summer (blue), and fall (magenta) seasons; (c) IPZ-2 (blue) at the Kingston Central and Kingston West intake. The red circle (1 km radius) is the IPZ-1. (For interpretation of the references to color in this figure legend, the reader isreferred to the web version of this article.)

10 20 30 40 50 60

5

10

15

20

25

30

5 cms−1

Nor

thin

g (k

m)

10 20 30 40 50 60

5

10

15

20

25

30

5 cms−1

Easting (km)

Nor

thin

g (k

m)

a

b

Fig. 13. Model velocity vectors averaged over June–August at (a) 12 m depth and(b) 20 m depth. The 20 m depth contour is shown in gray.

10 S. Paturi et al. / Journal of Great Lakes Research xxx (2011) xxx–xxx

Please cite this article as: Paturi, S., et al., Hydrodynamics of eastern Lake Ontario and the upper St. Lawrence River, J Great Lakes Res (2011),doi:10.1016/j.jglr.2011.09.008

Page 11: Journal of Great Lakes Research - Queen's University · 2020. 9. 15. · OOF Q31 Hydrodynamics of eastern Lake Ontario and the upper St. Lawrence River Q12 Shastri Paturi a, Leon

UNCO

RRECTED P

RO

OF

530 may be due to the inability of the open boundary to capture momen-531 tum transfer from the 400 km fetch to the southwest. ELCOM thus532 reasonably captures the dynamics of the flow regimes in the near-533 shore region.534 In agreement with Rao et al. (2009), the present hydrostatic535 model application with ~300 m grid resolution was unable to repro-536 duce observed high-frequency temperature fluctuations. These pro-537 cesses are likely associated with nonlinear/non-hydrostatic waves538 and shear instabilities (e.g., Boegman et al., 2003), and their resolu-539 tion requires a non-hydrostatic pressure solver and ~1–10 m grid res-540 olution. Such simulations are not presently feasible over seasonal541 timescales (e.g., Botelho and Imberger, 2007; Dorostkar et al., 2010).542 Future work will apply the validated ELCOM model to investigate543 the transport of river plumes and industrial and wastewater effluents544 within the model domain.

545 Acknowledgments

546 Funding was provided by the Ontario Ministry of the Environment547 Source Water Protection Technical Studies Program, Environment548 Canada, and Queen's University. Grid sensitivity runs were performed549 on computing facilities provided by the Canada Foundation for Inno-550 vation and the Ontario Innovation Trust. We thank J. Imberger at the551 Centre for Water and Research (CWR) for providing the ELCOM552 source code. Field observations and calibrations were coordinated553 by Bob Rowsell from Environment Canada. We thank Sean Watt and554 the CRCA for their support of this study.

555 References

556 Antenucci, J.P., Imberger, J., 2001. On internal waves near the high frequency limit in an557 enclosed basin. J. Geophys. Res (Oceans) 106 (C10), 22465–22474.558 Beletsky, D., Schwab, D.J., McCormick, M., 2006. Modelling the 1998–2003 summer cir-559 culation and thermal structure in Lake Michigan. J. Geophys. Res. 111, C10010.560 doi:10.1029/2005JC003222.561 Boegman, L., Rao, Yerubandi, R., 2010. Process oriented modeling of Lake Ontario hy-562 drodynamics. Proceedings 6th International Symposium on Environmental Hy-563 draulics, Jun. 23–25, Athens, Greece (6 pp.).564 Boegman, L., Imberger, J., Ivey, G.N., Antenucci, J.P., 2003. High-frequency internal565 waves in large stratified lakes. Limnol. Oceanogr. 48, 895–919.566 Botelho, D.A., Imberger, J., 2007. Dissolved oxygen response to wind–inflow interac-567 tions in a stratified reservoir. Limnol. Oceanogr. 52 (5), 2027–2052.568 Dallimore, C.J., Hodges, B.R., 2000. The Estuary and Lake Computer Model ELCOM. Sci-569 ence Manual, Centre for Water Research, University of Western Australia.570 Denison, N.F., 1908. Lake undulations. J. R. Astron. Soc. Can. 251.571 Dorostkar, A., Boegman, L., Diamessis, P., Pollard, A., 2010. Sensitivity of MITgcm to dif-572 ferent model parameters in application to Cayuga Lake. Proceedings 6th

573International Symposium on Environmental Hydraulics Jun. 23–25, Athens, Greece574(6 pp.).575Hall, E. 2008. Hydrodynamic modelling of Lake Ontario. M.Sc. Thesis. Department of576Civil Engineering. Queen's University.577Hamblin, P.F., 1982. On the free surface oscillations of Lake Ontario. Limnol. Oceanogr.57827, 1039–1049.579Hamblin, P.F., 1987. Meterological forcing and water level fluctuations on Lake Erie. J.580Great Lakes Res. 13, 436–453.581Hodges, B.R., 2000. Numerical techniques in CWR-ELCOM. Tech. Rep. WP 1422-BH,582Cent. for Water Res., Univ. of West Aust., Nedlands, West. Aust., Australia.583Hodges, B.R., Imberger, J., Saggio, A., Winters, K., 2000. Modelling basin-scale internal584waves in a stratified lake. Limnol. Oceanogr. 45, 1603–1620.585Huang, A., Rao, Y.R., Lu, Y., Zhao, J., 2010. Hydrodynamic modeling of Lake Ontario: an586intercomparison of three models. J. Geophys. Res. 115, C12076. doi:10.1029/5872010JC006269.588Laval, B., Hodges, B.R., Imberger, J., 2003. Reducing numerical diffusion effects with589pycnocline filter (2003). ASCE J. Hydraul. Eng. 129 (3), 215–224.590León, L.F., Imberger, J., Smith, R.E.H., Hecky, R.E., Lam, D.C., Schertzer, W.M., 2005.591Modeling as a tool for nutrient management in Lake Erie: a hydrodynamics592study. J. Great Lakes Res. 31 (Suppl. 2), 309–318.593Morillo, S., Imberger, J., Antenucci, J.P., Woods, P.F., 2008. Influence of wind and lake594morphometry on the interaction between two rivers entering a stratified lake. J.595Hydraul. Eng. ASCE 134, 1579–1589. doi:10.1061/(ASCE)0733-9429(2008)596134:11(1579).597National Building Code of Canada, 2005. National Research Council.598Ontario Ministry of the Environment, 2005. Assessment Report: Guidance Modules.599Source Water Implementation Group.600Paturi, S., Boegman, L., 2008. Delineation of intake protection zones in the CRCA juris-601diction — modeling approach. Technical Report Prepared for the Cataraqui Region602Conservation Authority.603Potok, A.J., Quinn, F.H., 1979. A hydraulic transient model of the upper St. Lawrence604River for water resource studies. : American Water Resources Association, Vol 15,605No. 6.. Water Resources Bulletin.606Rao, Y.R., Schwab, D.J., 2007. Transport and mixing between the coastal and offshore607waters in the Great Lakes: a review. J. Great Lakes Res. 33, 202–218.608Rao, Y.R., Marvin, C.H., Zhao, J., 2009. Application of a numerical model for circulation,609temperature and pollutant distribution in Hamilton Harbour. J. Great Lakes Res. 35,61061–73.611Schertzer, W.M., 1987. Heat balance and heat storage estimates for Lake Erie. J. Great612Lakes Res. 13 (4), 454–467.613Schwab, D.J., 1977. Internal free oscillations in Lake Ontario. Limnol. Oceanogr. 22,614700–708.615Shen, H.T., Yapa, P.D., Zhang, B., 1995. A simulation model for chemical spills in the616Upper St. Lawrence River. J. Great Lakes Res. 21 (4), 652–664.617Shore, J.A., 2009. Modelling the circulation and exchange of Kingston Basin and Lake618Ontario with FVCOM. Ocean Modell. 30, 106–114.619Thompson, A., Guo, Y., Moin, S., 2008. Uncertainty analysis of a two-dimensional hy-620drodynamic model. J. Great Lakes Res. 34, 472–484.621Tsanis, I.K., Murthy, C.R., 1990. Flow distribution in the St. Lawrence River system at622Wolfe Island, Kingston Basin, Lake Ontario. J. Great Lakes Res. 16 (3), 352–365.623Tsanis, I.K., Masse, A., Murthy, C.R., Miners, K., 1991. Summer circulation in the Kings-624ton Basin, Lake Ontario. J. Great Lakes Res. 17 (1), 57–73.625Vidal, J., Casamitjana, X., Colomer, J., Serra, T., 2005. The internal wave field in Sau res-626ervoir: observation and modeling of a third vertical mode. Limnol. Oceanogr. 50627(4), 1326–1333.

628

629

11S. Paturi et al. / Journal of Great Lakes Research xxx (2011) xxx–xxx

Please cite this article as: Paturi, S., et al., Hydrodynamics of eastern Lake Ontario and the upper St. Lawrence River, J Great Lakes Res (2011),doi:10.1016/j.jglr.2011.09.008