EVALUATING THE ROLE OF GROUNDWATER IN CIRCULATION, THERMAL STRUCTURE AND NUTRIENTALGAL DYNAMICS WITHIN A DEEP INLAND LAKE By Ammar Safaie Nematollahi A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Civil Engineering - Doctor of Philosophy 2017 ABSTRACT EVALUATING THE ROLE OF GROUNDWATER IN CIRCULATION, THERMAL STRUCTURE AND NUTRIENTALGAL DYNAMICS WITHIN A DEEP INLAND LAKE By Ammar Safaie Nematollahi Inland lakes readily respond to changes in external forcing and therefore serve as sentinels of climate change. Many parts of the world continue to experience declining groundwater levels due to anthropogenic activities such as high-capacity pumping for agriculture or decreases in natural recharge rates of aquifers while lake surface temperatures continue to rise and show a clear warming trend. The responses of individual lakes to these stressors could vary depending upon the positions of the lakes within the landscape and the nature of lake-groundwater interactions. Since temperature is a key driver that affects the structure and function of ecosystems including biological productivity, nutrient cycling and hypoxia, groundwater-fed lakes could be altered drastically due to declining groundwater contribution. Thus, it is crucial to understand the role of groundwater in biophysical processes and to determine what regime shifts may occur in the absence of lake-groundwater interactions. To address this question, extensive field datasets were collected in the Gull Lake, a deep, dimictic, groundwater-fed, inland lake in Michigan, with bottom cooling and strong stratification during summer. The lake supports diverse warm and cold water fisheries. Detailed three-dimensional hydrodynamic and temperature models of Gull Lake coupled to nutrient and algal dynamics were developed to study the effect of groundwater on physical, chemical, and biological processes in the lake. Coupled biophysical processes in the water column are closely linked to meteorological forcing. Therefore, meteorological forcing fields were carefully reconstructed from a network of weather station data, and were assessed using outputs from a mesoscale numerical weather forecasting model (WRF). A novel manifold method of reconstructing dynamically evolving spatial fields is presented for assimilating data from sensor networks in lake and watershed models. The manifold method has been developed based on the assumption that geophysical and meteorological data can be mapped onto an underlying differential manifold. A comparative evaluation of turbulence models was also conducted to improve descriptions of vertical mixing and thermal structure of the lake. The performance of the biophysical model was first evaluated against high-resolution in situ observations, including currents, lake levels, temperature, nutrients, dissolved oxygen, and chlorophyll data. After successfully applying the model to describe current conditions, the developed models were used to understand the responses of the lake ecosystem when the groundwater contribution is absent. Results suggest that groundwater-fed lakes have the ability to buffer seasonal water temperature variations in the hypolimnion, which helps them to withstand disturbances from surface-induced changes. However, groundwater depletion was accompanied by changes in the structure and function of lake ecosystems including lake level changes, rising water temperatures, increased growth rates of algae, oxygen depletion, early anoxia, reduction of light availability, and eutrophication. These results highlight the significant role played by groundwater in inland lakes and indicate that groundwater-dependent ecosystems tend to show greater resilience. In addition to providing insights into key biophysical processes in inland lakes, this study is expected to help strengthen management efforts to improve or maintain the resilience of lake ecosystems. Suggested citation: Safaie, A. (2017), “Evaluating the Role of Groundwater in Circulation, Thermal Structure, and Nutrient-Algal Dynamics within a Deep Inland Lake”, Ph.D. Dissertation, Michigan State University, East Lansing, Michigan, United States. Copyright by AMMAR SAFAIE NEMATOLLAHI 2017 iv To my family and my love v ACKNOWLEDGEMENTS This research was supported by a grant from the National Science Foundation, CyberSEES program (Award # 1331852). I would like to gratefully acknowledge my academic advisor, Prof. Mantha S. Phanikumar, for his enlightening guidance, enthusiastic encouragement, and continuous support throughout my doctoral program. I am thankful to my committee members, Prof. Farhad Jaberi, Prof. Elena Litchman, and Dr. Yadu Pokhrel for their time, suggestions, and feedback on my doctoral dissertation. Fieldwork and data collection would have not been possible without the assistance of Prof. Litchman and her crew, Pam Woodruff, Allyson Hutchens, and Kirill Shchapov, and I truly appreciate them all. I also thank Prof. Stephen K. Hamilton for sharing the inflow and outflow data for Gull Lake. Additionally, many thanks go to Mike Gallagher, Jim Allen, Andrew Fogiel, and Michael Nichols for their assistance with deployments of our instruments in Gull Lake. I am grateful to Prof. Hayder Radha and Dr. Chinh Dang for their collaboration and for introducing me to the manifold-based approach. I extend my appreciation to Prof. Shu-Guang Li and his research team, Dr. Huasheng Liao and my good friend Zachary K. Curtis, for their helpful discussions and great collaboration. My sincere thanks also go to my colleagues Tuan D. Nguyen, Guoting Kang, and Han Qiu for their fruitful discussions and assistance with field data collection. Most importantly, I would like to express my heartfelt thanks and gratitude to my family for their continuous support, dedication, encouragement, and prayers in every step of my life. Above all, I would like to express my sincerest gratitude to God for everything I have been given. vi TABLE OF CONTENTS LIST OF TABLES ......................................................................................................................... ix LIST OF FIGURES ........................................................................................................................ x CHAPTER 1 ................................................................................................................................... 1 1 Introduction .................................................................................................................................. 1 1.1 Problem description ............................................................................................................... 1 1.2 The study area ........................................................................................................................ 4 1.3 Summary of specific aims ...................................................................................................... 7 1.4 Expected benefits and significance ........................................................................................ 8 1.5 Dissertation structure ............................................................................................................. 9 CHAPTER 2 ................................................................................................................................. 11 2 Reconstruction of Geophysical and Meteorological Data in Integrated Earth System Models .... .................................................................................................................................................... 11 2.1 Introduction .......................................................................................................................... 11 2.2 Materials and methods ......................................................................................................... 15 2.2.1 Application of manifold methods for data assimilation in integrated Earth system models ......................................................................................................................................... 15 2.2.1.1 Manifold approach ................................................................................................... 15 2.2.1.2 Test case: Analytical function .................................................................................. 25 2.2.1.3 Assimilating meteorological data for improved lake circulation modeling: Lake Michigan .............................................................................................................................. 26 2.2.1.4 Assimilating geophysical data for improved lake circulation modeling: Gull Lake 29 2.2.2 Reconstruction of meteorological data in Gull Lake....................................................... 31 2.2.2.1 Land-based weather station network ....................................................................... 31 2.2.2.2 Mesoscale weather prediction model ....................................................................... 33 2.3 Results and Discussion ........................................................................................................ 35 2.3.1 Manifold method ............................................................................................................. 35 2.3.2 Observed and predicted meteorological data in Gull Lake ............................................. 41 2.4 Conclusions .......................................................................................................................... 46 CHAPTER 3 ................................................................................................................................. 48 3 Evaluating the Role of Groundwater in Circulation and Thermal Structure ............................. 48 3.1 Introduction .......................................................................................................................... 48 3.2 Materials and methods ......................................................................................................... 52 3.2.1 Observational data ........................................................................................................... 52 3.2.2 Numerical hydrodynamic model ..................................................................................... 54 3.2.3 Turbulent closure models ................................................................................................ 61 3.2.3.1 Mellor-Yamada 2.5 level turbulence model ............................................................ 61 3.2.3.2 The k turbulence model .................................................................................... 63 3.2.4 Light attenuation lengths ................................................................................................. 65 vii 3.2.5 Precipitation/Evaporation ................................................................................................ 66 3.2.6 Groundwater .................................................................................................................... 67 3.3 Results and discussion ......................................................................................................... 77 3.4 Conclusions .......................................................................................................................... 99 CHAPTER 4 ............................................................................................................................... 102 4 Evaluating the Role of Groundwater in Nutrient - Algal Dynamics ....................................... 102 4.1 Introduction ........................................................................................................................ 102 4.2 Materials and methods ....................................................................................................... 105 4.2.1 Field sampling and laboratory analyses ........................................................................ 105 4.2.2 Biophysical model ......................................................................................................... 106 4.3 Results and discussion ....................................................................................................... 117 4.4 Conclusions ........................................................................................................................ 126 CHAPTER 5 ............................................................................................................................... 128 5 Conclusions .............................................................................................................................. 128 APPENDIX ................................................................................................................................. 131 BIBLIOGRAPHY ....................................................................................................................... 134 viii LIST OF TABLES Table 2-1 Properties of the numerical grids used for the hydrodynamic models. ....................... 28 Table 2-2 Cross-validation results for the analytical function based on different sampling points selected randomly. ........................................................................................................................ 36 Table 2-3 RMSE values (m/s) of alongshore and cross-shore velocities in Lake Michigan for comparison of the manifold method with other standard methods used in limnology and oceanography. ............................................................................................................................... 37 Table 2-4 Cross-validation results for wind field over Lake Michigan. ...................................... 38 Table 2-5 RMSE values (m/s) of eastward and northward velocities in Gull Lake for comparison of the manifold method with other standard methods used in limnology and oceanography. ..... 41 Table 2-6 Cross-validation results for Gull Lake bathymetry...................................................... 41 Table 3-1 Details of instruments deployed in Gull Lake. ............................................................ 53 Table 3-2. Description of different models evaluated .................................................................. 60 Table 3-3. Averaged heat budget components of Gull Lake in the summers of 2014 and 2015. 80 Table 3-4. RMSE values (m/s) of eastward and northward velocities at ADCP locations. ......... 81 Table 3-5. RMSE values (m/s) of the observed and simulated vertical velocity profiles at ADCP locations for some selected time steps. ......................................................................................... 86 Table 3-6. Set of parameters used for the groundwater module. ................................................. 88 Table 4-1. Nutrient concentration ranges in Prairieville Creek and a pond lab reservoir well near Gull Lake from 2009-2014. ........................................................................................................ 106 Table 4-2 Parameter values used in the water quality model ..................................................... 116 ix LIST OF FIGURES Figure 1-1 Map of Gull Lake ......................................................................................................... 5 Figure 1-2 Gull Lake's watershed boundaries and digital elevation model (DEM) map. .............. 6 Figure 2-1 Some examples of manifolds (a) torus, (b) sphere, and (c) a two-dimensional crosssection of a six-dimensional Calabi-Yau manifold. ...................................................................... 16 Figure 2-2 Illustration of the proposed manifold approach for estimation of missing data at point P0. .................................................................................................................................................. 17 Figure 2-3 A tangent space created from the set of nearest points using (a) coordinates of selected neighborhoods or (b) Kernel regression. ...................................................................................... 19 Figure 2-4 Manifolds representing (a) bathymetry of Gull Lake and (b) wind components over Lake Michigan in three dimensional space. .................................................................................. 24 Figure 2-5 (a) Analytical function used to test the manifold method for interpolation of scattered data. Random sampling was used to generate scatter points as shown in figures (b, 30 points), (c, 60 points) and (d, 90 points) to reconstruct the function. ............................................................. 25 Figure 2-6 Locations of the ADCPs deployed during summer 2008 and weather stations surrounding Lake Michigan. ......................................................................................................... 27 Figure 2-7 (a) Bathymetry of Gull Lake. (b) Boat tracks generated during the sampling survey in Gull Lake. ..................................................................................................................................... 30 Figure 2-8 Selected weather stations surrounding Gull Lake. ..................................................... 32 Figure 2-9 Circular grid with a radius of 42 km used for the reconstruction of meteorological fields from the Land-based weather station network. ............................................................................. 33 Figure 2-10 Three nested simulation domains with resolutions of 30 km, 10 km, and 2 km used for the WRF model. ...................................................................................................................... 34 Figure 2-11 Performance of the manifold method evaluated using observed and simulated currents at different stations in Lake Michigan. Different number of nearest neighbors were used to reconstruct the wind field using the manifold method with kernel regression. ............................ 38 Figure 2-12 Comparison of simulated and observed vertically averaged currents at the location M in Lake Michigan. (a) Alongshore velocity (b) Cross-shore velocity .......................................... 39 Figure 2-13 Comparison of simulated (black lines) and observed (red lines) vertically averaged currents at the ADCP location in Gull Lake. (a) Eastward velocity and (b) Northward velocity 40 Figure 2-14 Comparisons of observed meteorological data at KBS LTER station with simulated results based on NCEP/NCAR data during April to mid-September of 2014 (a) air pressure (kPa), x (b) air temperature ( C ), (c) downward shortwave radiation (W/m2), (d) downward longwave radiation (W/m2), (e) relative humidity (%), (f) wind speed at 10-meter height (m/s), (g) eastward wind velocity (m/s), (h) northward wind velocity (m/s). .............................................................. 42 Figure 2-15 Scatter plots of WRF-simulated results versus observed meteorological forcing at KBS LTER during April to mid-September of 2014 (black dot) and 2015 (red dot): (a) air pressure (kPa), (b) air temperature ( C ), (c) downward shortwave radiation (W/m2), (d) downward longwave radiation (W/m2), (e) relative humidity(%), (f) wind speed at 10-meter height (m/s), (g) eastward wind velocity (m/s), (h) northward wind velocity (m/s)................................................ 43 Figure 2-16 Comparisons of averaged simulated forcing (WRF) and observed meteorological forcing (Stations) over Gull Lake during April to mid-September of 2014. ................................ 44 Figure 2-17 Comparisons of averaged simulated forcing and meteorological forcing over Gull Lake during April to mid-September of 2015............................................................................... 45 Figure 2-18 Comparisons of the simulated and observed wind speed in 2014 and 2015 at the KBS LTER station. ................................................................................................................................ 45 Figure 3-1 Map of Gull Lake showing locations of the in situ measurements. ........................... 53 Figure 3-2 Schematic of the thermistor chain .............................................................................. 54 Figure 3-3. The unstructured computational mesh created based on the bathymetry-based refinement algorithm ..................................................................................................................... 59 Figure 3-4. Quaternary geology of Lower Peninsula and Gull Lake. Figure adapted by author from Farrand [1982] ............................................................................................................................. 68 Figure 3-5. Bedrock Geology of the Lower Peninsula and the Gull Lake area. Figure adapted by author from Wilson [1987] ............................................................................................................ 70 Figure 3-6. (a) Groundwater flow map in the study area. (b) A cross-section of the aquifer along the longitudinal transect of the lake. ............................................................................................. 71 Figure 3-7. Time series of relatively constant bottom water temperatures (blue symbols) measured at a depth of 32 m between 2011 and 2015 compared to the surface water temperature and air temperature for the period 2011-2015 (black line). ...................................................................... 76 Figure 3-8. Water budget components in Gull Lake in (a) 2014 and (b) 2015 ............................ 77 Figure 3-9. Comparisons of water level fluctuations recorded by ADCPs (red line), simulated without groundwater (blue line), and simulated with groundwater (black and yellow lines) in (a) 2014 and (b) 2015. ........................................................................................................................ 78 Figure 3-10. Comparison of Gull Lake evaporation in 2014 computed by the mass transfer approach (method1) and the COARE algorithm (method2) ......................................................... 79 xi Figure 3-11. Comparisons of simulated water level fluctuations in Gull Lake using evaporation computed by the mass transfer approach (method1) and the COARE algorithm (method2). ...... 79 Figure 3-12. Comparison of simulated and observed vertically averaged salinity at HL location for year 2014. ................................................................................................................................ 80 Figure 3-13. Comparisons of observed and simulated vertically-averaged velocity profiles at ADCP locations in the summer of 2014. ...................................................................................... 82 Figure 3-14. Scatter plots of velocities (observed versus simulated for different models) for year 2014............................................................................................................................................... 83 Figure 3-15. Probability distributions of RMSE values based on observed and simulated vertical velocity profiles ............................................................................................................................ 84 Figure 3-16. Examples of comparisons of the observed and simulated vertical velocity profiles at ADCP locations in the summer of 2014. ...................................................................................... 85 Figure 3-17. Observed PAR data used to determine the parameters of the shortwave penetration model............................................................................................................................................. 87 Figure 3-18. Contour plots of the observed and simulated water temperature at the TC location in the summer of 2014. ..................................................................................................................... 88 Figure 3-19. Comparisons of simulated and observed time series of (a) surface water temperature (T0), and bottom temperature at (b) TC, (c) M14, and (d) S14 locations in 2014. ....................... 89 Figure 3-20. Taylor diagram which shows statistical comparisons of observed water temperature with simulated results of model 1 to 4. The statistics are normalized by the standard deviation of the observations ............................................................................................................................ 90 Figure 3-21. Comparisons of simulated vertical profiles of temperature with weekly Hydrolab data in 2014. .................................................................................................................................. 91 Figure 3-22. Comparisons of the simulated vertically averaged velocities and observations at ADCP locations in the summer of 2015. ...................................................................................... 92 Figure 3-23. Time series of wind, current and significant wave height at SS15 location. (a) Eastward wind speed (m/s), (b) comparison of observed and simulated eastward current speeds (m/s), (c) northward wind speed (m/s), (d) comparison of observed and simulated northward current speeds (m/s), (e) observed significant wave height (m). Gray areas represent examples of periods with significant wave activity during which the model could not adequately capture the observed currents. ......................................................................................................................... 93 Figure 3-24. Comparisons between observed (red lines) and simulated (black lines) power spectral densities at ADCP locations in 2014 and 2015. ............................................................................ 94 Figure 3-25. Examples of comparisons of the observed and simulated vertical velocity profiles at ADCP locations in the summer of 2015. ...................................................................................... 96 xii Figure 3-26. Contour plots of the observed and simulated water temperature at the thermistor chain location (TC) in the summer of 2015. ................................................................................. 97 Figure 3-27. Comparisons of simulated and observed time series of (a) surface water temperature at TC, and bottom temperature at (b) TC, (c) SS15, and (d) SD15 locations in 2015. ................. 98 Figure 3-28. Comparisons of simulated vertical profiles of temperature with weekly Hydrolab data in 2015. .................................................................................................................................. 99 Figure 4-1 Comparisons of simulated vertical profiles of dissolved oxygen (mg O2/L) with Hydrolab data in 2015. Simulated results of Model 1 (blue line) are in the absence of groundwater effects, and results of Model 2 (black line) are with groundwater effects. ................................. 118 Figure 4-2. Contour plots of (a) observed DO (mg O2/L), (b) simulated DO with groundwater effects (Model 2) and (c) simulated DO in the absence of groundwater effects (Model 1) in Gull Lake in 2015. .............................................................................................................................. 119 Figure 4-3. Comparisons of simulated vertical profiles of Chl-a (μg C/L) with Hydrolab data in 2015. Simulated results of Model 1 (blue line) are in the absence of groundwater effects, and results of Model 2 (black line) are with groundwater effects................................................................. 120 Figure 4-4. Contour plots of (a) observed Chl-a (μg C/L), (b) simulated Chl-a with groundwater effects (Model 2) and (c) simulated Chl-a in the absence of groundwater effects (Model 1) in Gull Lake in 2015. .............................................................................................................................. 121 Figure 4-5. Contour plots of (a) observed DO saturation (%), (b) simulated DO saturation with groundwater effects and (c) simulated DO saturations in the absence of groundwater effects in Gull Lake in 2015. .............................................................................................................................. 122 Figure 4-6. Comparisons of observed and simulated vertical profiles of dissolved nitrate (mg N/L) in 2015. Simulated results of Model 1 (blue line) are in the absence of groundwater effects, and results of Model 2 (black line) are with groundwater effects. .................................................... 123 Figure 4-7. Comparisons of observed and simulated vertical profiles of SRP (μg P/L) in 2015. Simulated results of Model 1 (blue line) are in the absence of groundwater effects, and results of Model 2 (black line) are with groundwater effects. .................................................................... 124 xiii CHAPTER 1 1 Introduction 1.1 Problem description Climate-induced changes in hydrodynamics of inland lakes are known to have physical, chemical, and biological effects on lake ecosystems. Wind speed, current shear, stratification and vertical mixing have significant effects on the cycling of nutrients, phytoplankton distribution, as well as on the growth rates of phytoplankton and benthic filter-feeders important for benthicpelagic trophic coupling [Denman and Gargett, 1983; Edwards et al., 2005; Rowe et al., 2015]. In addition, wind-mixing and thermal stratification have implications for regulating the supply of phytoplankton to zebra mussels [Boegman et al., 2008]. Changes in water temperature, thermal and mixing dynamics, strength and duration of stratification have impacts on primary production and associated food web, predation and growth rates of zooplankton and fish, nutrient supply, and deepwater anoxia [MacKay et al., 2009]. Hence, understanding coupled aquatic biophysical processes in lakes is critical to an improved understanding of lake ecosystems and their responses to climate change. Lake surface temperature is a key response indicator of climate change. Long-term air temperature and ice cover records suggest that freshwater ecosystems have experienced 1 temperature warming and shortening ice duration in response to the global warming over the past 150 years [Magnuson, 2000]. Remote sensing observations of inland water bodies since 1985 also provide evidence of significant warming trends in their surface temperatures to the extent that in some regions, such as the Great Lakes and Northern Europe, water temperature is rising more rapidly than regional air temperature [Schneider and Hook, 2010]. The lake warming has been reported even in the largest freshwater lakes around the word such as Lake Baikal [Hampton et al., 2008], Lake Superior [Austin, 2013], Lake Michigan [Brooks and Zastrow, 2002], Lake Erie [Burns et al., 2005], and Lake Tanganyika [O’Reilly et al., 2003]. These warming trends have resulted in longer summer stratifications, combined with higher summer water temperatures and a shorter winter-spring mixing period. Such changes in thermal stratifications have consequential impact upon the ecology of lakes. For instance, the changes could reduce primary productivity by limiting upwelling of nutrients from deep water [Brooks and Zastrow, 2002; O’Reilly et al., 2003; Chang et al., 2015], and change planktonic processes and community dynamics of other aquatic organisms further up in the food chain [Lane et al., 2008; MacKay et al., 2009; Rühland et al., 2015]. Surface fluxes are usually assumed to be the only important driving forces responsible for the aforementioned climate-induced changes. From a system dynamics point of view, if a system is composed of only a reinforcing loop, the closed loop of cause and effect would lead to instability via accelerating growth [Sterman, 2000]. Balancing loops, on the other hand, serve to resist attempted changes in order to maintain a balance and keep the system at a desired state. Identifying these reinforcing and balancing processes in a complex dynamic system like a freshwater ecosystem is crucial to better understand thermal structure, circulation, nutrients, and algal dynamics in inland lakes. It is known that primarily groundwater-fed rivers are buffered against 2 increasing seasonal temperature variation [Combes, 2003]. A shallow temperate lake fed by groundwater might also have a significant bottom heating in winter and bottom cooling in summer [Kettle et al., 2012]. This groundwater effect may act as a balancing loop within a lake system. If this hypothesis is true, then groundwater can be introduced as a key contributor to offset warming and anoxia in the groundwater-fed lakes. However, long-term groundwater depletion has been reported in many different regions of the world [Rodell et al., 2009; Wada et al., 2010; Konikow, 2013; Joodaki et al., 2014; Pokhrel et al., 2015; Dalin et al., 2017]. Therefore, groundwater dependent ecosystems are likely to be under dual pressure from surface temperature warming and groundwater depletion in response to natural climate variability and human activities. Thus, in order to protect ecosystems of groundwater-fed lakes, we need to understand the role of groundwater in their biophysical processes, and determine what regime shifts may occur in their ecosystems in the absence of lake-groundwater interactions. There are a variety of numerical ocean models, such as POM [Blumberg, and Mellor, 1987] ROMS [Haidvogel and Beckmann, 1999], and FVCOM [Chen et al., 2003b, 2006] that can be used to predict and obtain a better understanding of aquatic biophysical processes. Performance of these mechanistic models is highly dependent on accuracy of surface forcing fields (i.e. wind stress and heat flux). These forcing fields can be obtained from observational data, the output of a weather forecast model, or a combination thereof [Xue et al., 2015]. In the Great Lakes region, real-time and historic meteorological data are available from buoys and monitoring stations across the lakes. However, realistic meteorological forcing fields (reconstructed using point observations) are still needed and the hydrodynamic results could be enhanced by using an improved algorithm for reconstruction of meteorological fields based on data from a network of weather stations [Safaie et al., 2016] or utilizing a weather forecast model [Xue et al., 2015]. In contrast, for 3 relatively small inland lakes nestled within the landscape, meteorological forcing data are not always available and intraregional data have to be compiled from local weather stations (e.g., from airports) several kilometers away from the lake [Hondzo and Stefan, 1993; Markfort et al., 2010]. Moreover, meteorological data from a single station near the lake may not be representative of the spatial distribution of meteorological fields around the lake and may produce unreasonable results [Rueda et al., 2005]. 1.2 The study area To evaluate the role of groundwater, we used field datasets and a coupled biophysical model of Gull Lake, a relatively small (8 km2 surface area) but deep (34 m maximum depth) clear water lake in the in Kalamazoo County in southwestern Michigan. Gull Lake is a hardwater, groundwater-fed lake, representative of many other inland lakes in the Great Lakes region. This lake is an attractive recreational site that supports diverse warm and cold water fisheries. Over 70 species of fish with a seasonal pattern in their predator dynamics are reported in Gull Lake [Lane, 1979]. The presence of a Michigan State University research field station (the W.K. Kellogg Biological station), located on shores of Gull Lake, has resulted in long term biological and limnological research. In 1941, a bathymetric map of the lake was created to aid fisheries research. The earliest limnological study of Gull Lake by Perry and Brown [1942] reported the thermal stratification in the lake. The lake is usually stratified from May to early November. Gull Lake was of great scientific interest following the zebra mussel invasion in 1994, and the subsequent rapid increase of Microcystis biomass after the invasion [Wilson and Sarnelle, 2002; Sarnelle et al., 2005; Bruesewitz et al., 2009; Horst et al., 2014; White et al., 2015]. Gull Lake nitrification and denitrification rates in littoral sediments are relatively high likely due to the presence of zebra mussels [Bruesewitz et al., 2009]. Observations from 1965 to 1975 showed a high growth rate of 4 algae in Gull Lake which was controlled by phosphorus loading into the lake. Long term phosphorus concentration of Gull Lake has been stable since 1994, and currently the water quality in the lake is reported to be in good condition. In summers, however, some concerns have been raised about blue-green algal blooms [White and Hamilton, 2014]. Prairieville Creek provides the main surface inflow into Gull Lake from the north. This tributary had a small annual flow rate of 0.19 (m3/s) in 2014. The lake has a single outflow southward to the Gull Creek (Figure 1-1). The flow rate of the outflow and the lake level are controlled by a sluice-gate dam. Gull Lake also receives water from three small lakes: Little Long, Wintergreen, and Miller Lakes. All these lakes have glacial kettle type basins, and the soil in this area is classified as glacial outwash sand and gravel, and postglacial alluvium. Figure 1-1 Map of Gull Lake 5 Bedrock type divides Gull Lake into two bedrock aquifer systems. The northern half of the lake is underlain by a permeable Marshall formation, and the remaining part consists of Coldwater shale with a relatively low permeability. The topographic elevation around the lake ranges between 237 to 305 meters with a downward slope toward the south/southeast (Figure 1-2). It has been reported that, other than small tributaries, inflow of Gull Lake is provided by groundwater discharge through springs from the bottom of the lake [Perry and Brown, 1942; Moss, 1972]. Tague [1977] estimated the water budget of Gull Lake in 1974, with 40% of its water from groundwater inflow, 25% from direct precipitation onto the lake surface, and 35% from stream inflows. Figure 1-2 Gull Lake's watershed boundaries and digital elevation model (DEM) map. During the summer stratification period, Gull Lake is alkaline with average pH values of 9 and 8.2 in the epilimnion and hypolimnion, respectively. The level of alkalinity reflects the 6 interaction of the lake with groundwater. Kinsman-Costello et al. [2015] reported that the groundwater contribution to the water budget of the lake was as high as 90% in late summer of 2010. This percentage of the groundwater influence was calculated from magnesium ion (Mg2+) concentrations in shallow (<2m deep) waters as a conservative tracer for groundwater, assuming that groundwater and precipitation are the only source of dissolved Mg2+ in the lake. In inland shallow lakes, heat flux exchange between water and lake sediment needs to be taken into consideration for vertical thermal diffusivity analyses [Hondzo et al., 1991]. It was suggested that heat exchange between water and lake sediments in inland shallow lakes needs to be taken into consideration for vertical thermal diffusivity analyses [Hondzo and Stefan, 1993]. In most hydrodynamic models, however, the role of groundwater exchange in the energy budget and seasonal stratification are assumed to be negligible. 1.3 Summary of specific aims The major objective of this study was to evaluate the role of groundwater in physical, chemical, and biological processes in Gull Lake which is a typical representation of many other groundwater-fed inland lakes in Michigan. Despite decades of water quality monitoring and biological research in Gull Lake, a numerical model of the lake has not been developed to date. A major motivation of this work is to develop a hydrodynamic model of Gull Lake coupled to nutrient (N, P) and algal dynamics. The working hypothesis here is that groundwater flow through Gull Lake causes significant bottom cooling in summer, to an extent that it has an important role in offsetting hypolimnetic warming and anoxia in the lake. Groundwater also has an effect on temperature, and it is well-know that phytoplankton and algae growth rates are dependent on temperature. This work is an attempt to quantify contributions of groundwater in a lake ecosystem by addressing the following research questions: 7 1. What is the role of groundwater in circulation and thermal structure within Gull Lake? Aim 1: Evaluate the role of groundwater in circulation and thermal structure. Aim 2: Quantify the impact of ignoring groundwater contribution on temperature. Aim 3: Quantify the percentage contribution of groundwater to the water budget of the lake. 2. What is the role of groundwater in water quality of Gull Lake? Aim 1: Understand the effect of groundwater on dissolved oxygen. Aim 2: Understand the effect of groundwater on nutrient-algal dynamics. Aim 3: Understand the impact of ignoring groundwater contribution on nutrients and algae. 1.4 Expected benefits and significance The aim of this research is to combine a coupled biophysical model and field observations to provide improved understanding of physical and biological behaviors of lakes. This research highlights the importance of algal blooms in the context of climate change and how groundwater tends to slow down the effects of the climate change. Insights gained from the developed complex mechanistic model will advance the understanding of the response of dissolved oxygen in inland lakes to the warming trend of air temperature throughout the world. Another aspect of this study is to document a systematic approach to modeling hydrodynamics and water quality of inland lakes that can be used for other fresh water bodies. These outcomes could strengthen management efforts to decrease nutrient loads, control hypolimnetic oxygenation, and monitor human activities that affect groundwater – surface water interactions. 8 1.5 Dissertation structure A coupled three-dimensional hydrodynamic and water quality model of Gull Lake was developed to investigate the effect of groundwater on physical, chemical, and biological processes in the lake. The model was refined and assessed through the following steps: 1. Collecting accurate bathymetry of the lake in order to improve the performance of the numerical model. 2. Developing a novel manifold-based method to assimilate geophysical and meteorological data in integrated Earth system models for better representations of the meteorological fields and bathymetry. 3. Evaluating the performance of the model for different forcings including a mesoscale weather forecast model. 4. Refinement of the unstructured numerical mesh horizontally and vertically using a bathymetry-based refinement algorithm to improve the accuracy of simulated circulation and thermal structure. 5. A comparative study of turbulence models to identify superior formulations and to further improve descriptions of vertical mixing and thermal structure of the lake. 6. Utilizing high-resolution in situ observations to evaluate the performance of the models in describing coupled biophysical processes. The developed models were then used to predict the responses of the lake ecosystem caused by disconnection of the lake from groundwater. This doctoral dissertation is divided into the following five main chapters: Chapter 1. Introduction 9 Chapter 2. Reconstruction of geophysical and meteorological data in integrated Earth system models Chapter 3. Evaluating the role of groundwater in circulation and thermal structure Chapter 4. Evaluating the role of groundwater in nutrients, and algal dynamics Chapter 5. Conclusions 10 CHAPTER 2 2 Reconstruction of Geophysical and Meteorological Data in Integrated Earth System Models 2.1 Introduction Performance of integrated Earth system models relies upon accuracy of meteorological and geophysical data. In situ observations generally have sparse and inhomogeneous distribution in space and time, and it is often infeasible to accurately reconstruct the true field from the data. However, more information about the structure of the field and its evolution, allows for better approximations [Barth et al., 2008]. For instance, currents in large lakes such as the Laurentian Great Lakes are mostly controlled by wind. Therefore, by improving the representation of wind fields in models of lake circulation, we expect to describe coupled biophysical processes in lakes more accurately. For example, Safaie et al. [2016] demonstrated that improved representation of meteorological fields based on natural neighbor interpolation of weather station data produced superior results for currents and bacterial concentrations relative to similar results based on a nearest neighbor interpolation of the same data. Accurate representation of geophysical features such as topography and bathymetry is also important in Earth system models and their components, 11 and model performance depends on the interpolation method used to assign the topographic information over a numerical mesh in processed-based models. In order to assimilate observations into the models, and estimate variables at unsampled locations and/or times, it is crucial to use a suitable interpolation method. Various deterministic [e.g., nearest neighbor, natural neighbor, inverse distance weighting (IDW), spline, polynomial] and geostatistical (e.g. kriging) interpolation methods have been developed to generate spatial fields. There have been numerous efforts to compare different spatial interpolation methods in order to identify the best method for a given model application. Yan et al. [2014] compared different interpolation methods, including IDW, global polynomial interpolation, local polynomial interpolation, radial basis functions, ordinary kriging (OK), simple kriging (SK), universal kriging (UK), and co-kriging (CK) to determine the water/land boundary point elevation based on in situ water level data from 14 control stations in Dongting Lake. They used a cross-validation method to select the optimal method, which was found to be the OK method. Merwade [2009] studied the effect of spatial trend on interpolation of river bathymetry, and compared the performance of different interpolation methods. The number of measurements and their spatial arrangement, as well as channel morphology and geology were found to influence the accuracy of the interpolation results [Merwade, 2009]. Due to the effects of these and other factors on the performance of various methods, comparisons of different spatial interpolation methods could not point out the best universal interpolation method [Li and Heap, 2008; Šiljeg et al., 2015]. Many studies, such as aforementioned works, have used cross-validation for assessing the performance of the interpolation methods. In this method, a subset of the original dataset is withheld to be used later for validating the interpolated field constructed from the rest of the observational data. Mean error (ME), root mean square error (RMSE) and the coefficient of 12 determination (R2) are commonly used to evaluate the performance of each interpolation method [Suparta and Rahman, 2016]. However, every problem has a unique method of interpolation that works best for a given distribution of observations and the intended use of the interpolated data. Density of a sensor network, spatial variability of the variable of interest and its distribution, and observational errors, all influence the accuracy of the interpolated field [MacEachren and Davidson, 1987]. For example, Luo et al. [2008] compared seven spatial interpolation techniques to identify which method produced the best estimation of the wind speed data recorded across England and Wales. Their study showed that kriging is the best method, and that the thin plate spline method had higher ME and RMSE values. However, in Suparta and Rahman [2016] the performance of the thin plate spline interpolation based on the RMSE and R2 values was found to be better than kriging for less dense data points over the selected interpolation surface. Therefore, comparing interpolation methods using the cross-validation method without considering the data structure and the purpose of interpolation is not guaranteed to produce the best representation of the underlying data. In this study, a novel manifold method is proposed to assimilate different types of spatiotemporal data in integrated Earth system models based on the hypothesis that an environmental dataset (including independent variables such as longitude, latitude, and time, and the measured variables of interest) can be mapped onto an underlying differential manifold. Working directly in the high dimensional space generally involves dealing with complex algorithms. Modeling the high dimensional data using manifolds with fewer degrees of freedom has captured a great deal of attention recently [Zhang et al., 2016a]. The use of low-dimensional manifolds not only reduces computational load for further processing, but also helps visualize the 13 entire dataset, which is an important initial step to make sense of the data before proceeding with more goal-directed modeling and analyses [Ma et al., 2011]. The problem of non-linear dimensionality reduction for a set of high dimensional data points is known as manifold learning. Examples of early works for non-linear dimensionality reduction include Isomap [Tenenbaum et al., 2000], local linear embedding (LLE) [Roweis and Saul, 2000], and Eigenmaps [Belkin and Niyogi, 2003], which have been used to learn the true underlying lowdimensional manifold structure of the data. Since then, the manifold model has been exploited extensively in numerous applications such as face recognition, action classification, segmentation, image denoising, image/video super-resolution, and multi-scale image analysis [Carin et al., 2011; Dang et al., 2013; Dang and Radha, 2014]. Most of these manifold learning methods have been inspired by linear techniques, mainly based on the assumption that non-linear manifolds can be approximated by locally linear parts [Mordohai and Medioni, 2010]. Two pioneering works in this area are the Isomap approach [Tenenbaum et al., 2000] and the LLE algorithm [Roweis and Saul, 2000]. The Isomap algorithm aims to preserve the geodesic distance among points from the input dataset. On the other hand, the LLE algorithm targets the local linear geometry of neighbors in a manifold. Numerous works on manifold learning have been further developed. A comprehensive review of prior works can be found in van der Maaten et al. [2009]. In this chapter, the effectiveness of the presented manifold algorithm is evaluated through assimilation of geophysical and meteorological data in lake models (section 2.1), although the methods described are general and can be used in many other areas of computational geosciences. We first apply the proposed method to reconstruct wind fields (time-varying vector fields) over Lake Michigan. Since currents in Lake Michigan are primarily driven by wind, we expect to improve the simulation of hydrodynamic and biophysical variables of interest by improving the 14 wind fields. Instead of relying on the cross-validation of interpolated wind data, however, we use a well-tested hydrodynamic model of Lake Michigan and compare current measurements with simulated currents to test the interpolation methods. The manifold methods are used to reconstruct meteorological data, which are available from buoys and monitoring stations across the lakes, to improved simulation of circulation in Lake Michigan. Then the method is applied to assimilate bathymetry data as a scalar field for use in a hydrodynamic model of Gull Lake. For small inland lakes nestled within the landscape, a nearby network of meteorological stations are not always available. In addition, meteorological data from a single station near the lake or from local weather stations several kilometers away from the lake may not be representative of the spatial distribution of meteorological fields around the lake. Therefore, we used a mesoscale weather prediction model along with data from a network of land-based weather stations to assess the accuracy of reconstructed forcing over Gull Lake (section 2.2). The predicted and observed weather data will be used in Chapter 3 to run the hydrodynamic model of Gull Lake. 2.2 Materials and methods 2.2.1 Application of manifold methods for data assimilation in integrated Earth system models 2.2.1.1 Manifold approach A manifold ( M ) is an n-dimensional topological space such that each point of M and its neighborhood can be approximated by a small flat piece in the Euclidean space, n . We can think of a manifold as a set of low-dimensional curves and surfaces within higher dimension Euclidean spaces [Guillemin and Pollack, 2010]. Some typical examples of manifolds are smooth surfaces, such as a torus (Figure 2-1a) or a sphere (Figure 2-1b), where each point and its neighborhood can 15 be approximated by a small flat linear-subspace within the three-dimensional Euclidean space. Another example of a manifold in a high dimensional space is a Calabi-Yau manifold which has found important applications in theoretical physics (e.g. superstring theory). Figure 2-1c shows a two-dimensional cross-section of a six-dimensional Calabi-Yau manifold. Surfaces of all these three manifolds are not a Euclidean space. The laws of the Euclidean geometry, however, are valid locally. Figure 2-1 Some examples of manifolds (a) torus, (b) sphere, and (c) a two-dimensional cross-section of a six-dimensional Calabi-Yau manifold. Based on Einstein’s theory of relativity, physical events are located on the continuum (manifold) of space-time. Therefore, station locations and times of observations form a space-time manifold viewed as a four-dimensional vector space. One way to handle spatiotemporal interpolation problems, inspired by this concept, is to integrate space and time simultaneously [Li and Revesz, 2004]. An underlying assumption behind this approach is that time and space dimensions can be treated as equally important [Li et al., 2014a]. In order to add time as another dimension of space, time values are needed to be scaled for a spatiotemporal dataset by a scaling speed [Schwab and Beletsky, 1998; Li et al., 2014a]. For a point measurement, we can then define a four-vector P (ct , x ) where c is a time scale, t is the time coordinate and x is a three- 16 dimensional vector space. We assume that the set of high-dimensional data points P (and the estimated data points P0 ) belongs to a differential manifold M , which may be curved and have a complicated topology, but the neighborhood of each point is approximately similar to a small piece of Euclidean space (resembles D ). Since a traditional distance measure is built upon the geometry of Euclidean space, we adapt the calculation to a neighborhood or a small region of the assumed manifold. An example of a one-dimensional curve in Figure 2-2 illustrates the general idea of the manifold estimation approach. The set of points P in Figure 2-2 includes sample data points where we have measured data as well as a point P0 where data are missing. For example, in the context of the wind field data, one full measurement (or data point) includes five components: time, longitude, latitude, wind speed, and wind direction. The partially missing data point may contain known components (time, longitude, latitude) and unknown or missing components (wind speed and wind direction). Figure 2-2 Illustration of the proposed manifold approach for estimation of missing data at point P0. Suppose that it is desired to estimate the wind field for a data point P0 n (n dimension of vector space) from a set of training data points that belong to a manifold M . The space/time coordinates of the point (the independent variables) are known. However, the data (the dependent 17 variable) are missing. We denote P0 P0 n P0 as the data point using the superscript  to denote the independent variables and the superscript  to denote the dependent variable which is the missing component of interest here. P0 P0 n (n is the sub-vector of the known components, and n) is the corresponding sub-vector field (e.g., wind vector) for the missing n component where P0 V (u , v ) and u and v are the orthogonal components of the wind ( V ). The training data points, for example P components Pi n Pi Pi n {P1 , P2 ,..., P7 } in Figure 2-2, also include the two , but there is no missing component here since both dependent and independent variables are assumed to be known at the nearby stations. Given a point P0 n , the algorithm locates a set of nearest points to P0 based on the distances d ( Pi , P0 ) between pairs of points Pi and P0 . In order to determine local neighbors of P0 , we can calculate the distances between P0 and either all other points within a fixed radius ε, or all of its k nearest neighbors [Tenenbaum et al., 2000]. Then, a tangent space (linear subspace) of the manifold M at the point P0 is created from the set of nearest points (Figure 2-3a), denoted byTP0 (M ) T T where T ,T denote the tangent spaces for the independent and dependent variables in the data at P0 and P0 .  P0  Finally, the point P0       P0  n will be located as the closest point that belongs to that tangent space. 18 Figure 2-3 A tangent space created from the set of nearest points using (a) coordinates of selected neighborhoods or (b) Kernel regression. The tangent space approximation is describe as follow. Consider a smooth n-dimensional manifold M embedded in a D-dimensional Euclidean space. To understand the local geometry of the surface f ( x ) near a point x  n , we consider the first-order Taylor series expansion of the surface: f  x   f ( x)  where J f  x   f  x  x D n  x  x   O( x  x 2 )  f ( x)  J f  x  x   O( x  x ) 2 (1) is the Jacobian matrix of f at the point x . If the components of f ( x ) can be defined as: f ( x)   f1 ( x), f 2 ( x), f 3 ( x) f D ( x)  and x   x1 , x2 , x3 T be written as: 19 xn  , then the Jacobian can T  f1  x  1 J f  x     f D  x1 f1  xn     f D  xn  (2) To understand the local shape of the surface in Eq. (1), we seek to determine the space  x  x  , such that as we move away from x , the value of the function does not change to within first order. This is equivalent to finding the space T such that: T  x  x  J f  x  x  x   0 (3) This space is the tangent space to the surface at point x and is the right null space of the Jacobian matrix J f ( x ) . The space orthogonal to the tangent space is the row space of the Jacobian and orthogonal representations of these spaces can be obtained from SVD:  0 V T  J f  U U     T   0 0 V  (4) where V spans the tangent space (right null space) and V spans the row space. Alternatively, in Eq. (4), the right null space of J f is the columns of V corresponding to zero singular values. Therefore, the tangent space of the manifold M at y  f ( x ) is: T ( M )  span  J f  x   (5) From a practical computation point of view, given a set of sample points y  { y1, y2 , y3 , ym} , the tangent space can be directly estimated using SVD. If C m denotes the local covariance matrix: 20 Cm  1 m yi yiT  U U T  m i 1 where U  u1, u2 , u3 , (6) uD  and   diag 1, 2 , 3 D  denote the eigenvector and eigenvalue matrices respectively, then the optimal (in a least-squares sense) n-dimensional linear subspace is the span of the n-largest eigenvectors in U: T (M )  span u1 , u2 , u3 , un  (7) Additional details and other methods of estimating the tangent space are described in Dang et al. [2014]. Now by having the tangent space, we can use the Euclidean distance of an orthogonal projection from a point to the tangent space to represent the closest distance between that point and the tangent space. Since a tangent space is a linear space (or affine space in a more general case), one point can orthogonally project into that space. The question is how to define neighbors for each data point? The underlying idea is how to define similarity distance among the training data points, and then the overall similarity matrix. Several methods have been considered in the past, such as k-nearest neighbors [Press, 2007], ϵ-ball method [Allard et al., 2012] or the use of sparse representation theory [Dang et al., 2014; Dang and Radha, 2015]. The estimation of P0 is performed using the following steps: 1. Given a set of neighboring points, estimate the tangent space T at the point of interest, P0 : Details of the method for creating a tangent space from a set of data points were described above. One simple method is to create a tangent space using singular value decomposition (SVD, Press [2007]). By way of an example in Figure 2-2, a tangent space (line b) is created for P0 from a set of its neighboring points ( P5 and P6 ). This tangent space at P0 21 M is denoted by T . 2. Find the orthogonal projection of P0 onto the tangent space: The closest point P T to the given point P0 is located at the intersection of the line b and the line perpendicular to it which passes through the point P0 . P which is a projection of P0 onto the subspace T can be represented as an approximation of point P0 . The orthogonal projection of vector point P0 in a high-dimensional space onto a low-dimensional vector subspace is given by: T A AT A ( P0 ) where A P0 and T T 1 D n AT P0 AA P0 (8) is a full rank matrix containing the set of points on the tangent space of ( P0 ) denotes the projection of P0 onto the subspace T . This projection is derived from the solution of the normal equation AT Ax AT P0 which is equivalent to the associated least squares solution of Ax P0 . Due to the difficulty associated with inverting a general matrix that may be singular or non-square depending on the number of neighboring points selected in the manifold method, the problem (1) can be posed as a minimization problem in which the MoorePenrose pseudoinverse A+ [Golub and Van Loan, 2013] of the original matrix A is used. The pseudoinverse A generalizes the concept of matrix inverse and arises in the minimum norm (that is, approximate as opposed to exact) or best-fit (in a least squares sense) solution to a system of linear equations. The problem minimize Ax  P0 x  2  has the solution: x  A P0 . The pseudoinverse can be computed using SVD as follows: if A U V T , where U , V denote unitary matrices and  is a diagonal matrix containing the singular values of A , then A V U T . The function pinv was used to compute the pseudoinverse in MATLAB. 22 3. Find a linear representation coefficient vector α of that projection onto the tangent space: This coefficient is calculated by solving the following equation: T ( P0 ) A (9) 4. Estimate the missing components of the point P0 ( P0 ): The last step is finding a point on the subspace T that is closest (in norm) to the point P0. In order to do that, T is projected using the projection coefficient α computed in step 3: P0 (10) T . The result of this projection is the closest point to P0 that belongs to its subspace. In this algorithm, high-dimensional coordinates of selected neighborhoods on the manifold are projected to a low-dimensional subspace. An alternative to this approach is to use kernel regression to assign a weight to each neighbor based on the distance from P0 (Figure 2-3b). A weight for each selected neighborhood can be computed using the following Gaussian kernel function: Wi  e   P P  i o 2 2 2  ,   var d  Pi  , Po   (11) Examples of manifolds representing geophysical (bathymetry) and meteorological (wind) data are shown in Figures 2-4(a) and (b). These figures support the assumption that the manifold can be considered as being linear locally, but with complicated topology overall. 23 Figure 2-4 Manifolds representing (a) bathymetry of Gull Lake and (b) wind components over Lake Michigan in three dimensional space. 24 2.2.1.2 Test case: Analytical function Before applying the manifold method to reconstruct complex geophysical and meteorological data, we first evaluate the effectiveness of the method in reproducing an analytical function, since errors can be computed relative to the known function values; therefore, the F7 function suggested by Lazzaro and Montefusco [2002] and Renka and Brown [1999] is used: F 7( x, y) 2cos(10 x) sin(10 y) (12) sin(10 x y) where the domain of F7 is restricted to 0 x 1 and 0 y 1 (Figure 2-5a). Three sets of sparse random points from a normal distribution were generated in the domain with numbers of sampling points of 30, 60, and 90. The F7 function was sampled randomly as shown in Figure 2-5b. Figure 2-5 (a) Analytical function used to test the manifold method for interpolation of scattered data. Random sampling was used to generate scatter points as shown in figures (b, 30 points), (c, 60 points) and (d, 90 points) to reconstruct the function. 25 The manifold method was tested by withholding one point at a time and estimating its associated value from the remaining points using the manifold method, in addition to other methods such as the natural neighbor, nearest neighbor, and IDW interpolations. Since known components of the scatter points are located in the two-dimensional X-Y plane, at least two neighboring points are needed to form a tangent space for the manifold method. Therefore, for simplicity, only two nearest neighbors are used in both manifold and IDW interpolation methods. 2.2.1.3 Assimilating meteorological data for improved lake circulation modeling: Lake Michigan The proposed method was first applied for the reconstruction of wind fields (time-varying vector fields) over Lake Michigan. Hourly wind speed and direction data during April-September 2008 were obtained from the National Data Buoy Center (NDBC) weather stations surrounding the lake (Figure 2-6). The wind measurements were adjusted to a 10 m anemometer height using the profile methods described in Schwab (1987). Since the aerodynamic roughness over the lake is much lower compared to its counterpart over the land, an empirical overland-overlake adjustment was applied to the wind speeds recorded by overland stations [Schwab and Beletsky, 1998]. The datasets of wind speed and direction were converted to two coordinates in the Cartesian coordinate system (𝑥 and 𝑦 directions). Instead of using the cross-validation method to evaluate the interpolated wind data, results from the hydrodynamic model of the lake were compared with current measurements to test the applied method. To this end, a well-tested three-dimensional hydrodynamic model of the lake [Safaie et al., 2016] was used. The model was based on the unstructured grid Finite Volume Community Ocean Model (FVCOM; Chen et al. [2006]) which was successfully used in the past in ocean [Li et al., 2014b], lake [Nguyen et al., 2014] and river [Anderson and Phanikumar, 2011] 26 modeling. Details of the unstructured mesh used in the hydrodynamic model are presented in Table 2-1. Figure 2-6 Locations of the ADCPs deployed during summer 2008 and weather stations surrounding Lake Michigan. Wind fields from April to September 2008 were reconstructed at the locations of nodes in the numerical mesh. Other hourly meteorological observations related to heat flux fields, including air temperature, cloud cover, dew point, long-wave solar radiation, short-wave solar radiation, and relative humidity, obtained from the National Climatic Data Center (NCDC) and NDBC stations, were interpolated over the computational grid using a smoothed natural neighbor method with a 27 smoothing radius of 30 km. Air pressure was assumed to be constant (105 Pa) through the course of the study and a constant startup water temperature with a value of 2.5 oC was used in the model. Table 2-1 Properties of the numerical grids used for the hydrodynamic models. Site Grid Element Grid #Vertical # Nodes # Elements Classification shape Resolution layers Lake Michigan Unstructured Triangle Gull Lake Triangle Unstructured 4m -5 km 12,684 8-100m 5,132 23,602 20 9,361 20 The overlake dew points were estimated from overland observations using an empirical formula described in [Schwab and Beletsky, 1998]. Air temperature and cloud cover were used to estimate long-wave solar radiation [Parkinson and Washington, 1979] and short-wave solar radiation was modeled using the clear-sky value and cloud cover [Nguyen et al., 2014]. Six arcsecond bathymetric data obtained from the NOAA National Geophysical Data Center (NGDC) combined with two-meter resolution LIDAR data along the Indiana coast from the National Oceanic and Atmospheric Administration (NOAA) were interpolated to the numerical mesh using the natural neighbor method [Safaie et al., 2016]. Three bottom-mounted, upward-looking Acoustic Doppler Current Profilers (ADCPs) were deployed at stations M, BB and S (Figure 2-6) in southern Lake Michigan from early June to late August 2008 to measure nearshore currents for model testing [Thupaki et al., 2013; Safaie et al., 2016]. The hydrodynamic model was run from April to August 2008 to have a two-month spin-up period. Evaluation of the manifold method was carried out by comparing the simulated currents with data collected by the ADCPs. Comparisons between simulated and observed currents can be improved by identifying an optimal set of parameters in the manifold method. These parameters 28 include: an optimum number of the nearest neighbors to create a tangent space, the time scale c, and parameters of the Gaussian kernel function. In addition, the method used for creating a tangent space from a set of data points can be changed to improve the agreement between simulated and observed currents. The manifold method for the reconstruction of wind fields was directly applied to reconstruct the other six scalar observations, including air temperatures, cloud cover, dew point, relative humidity, shortwave and longwave solar radiation, to calculate the heat flux fields. This time, however, 𝑃𝑣 is a scalar, rather than a vector. 2.2.1.4 Assimilating geophysical data for improved lake circulation modeling: Gull Lake In the second example, the bathymetry of Gull Lake was reconstructed using a manifold method. The lake bathymetry data were collected using a SonTek RiverSurveyor M9 system. The M9 system has an Acoustic Doppler Profiler (ADP) with two sets of four profiling beams and one vertical acoustic beam (0.5-MHz echo-sounder) for river discharge measurements and bathymetric surveys. The system was equipped with differential GPS with sub-meter precision and mounted on a SonTeck hydroboard to avoid high pitch and roll angles. The vertical acoustic beam has a range of 0.2 m to 80 m with an accuracy of 1% and a resolution of 0.001 m. The bathymetry survey was performed in four days (June 9 – June 12, 2015) by collecting data along longitudinal and transverse transects of the lake with an approximate interval of 200 m between each transect pair and sampling interval of 0.2 m- 2 m along the transects depending on the boat speed (Figure 2-7). 29 Figure 2-7 (a) Bathymetry of Gull Lake. (b) Boat tracks generated during the sampling survey in Gull Lake. In order to assimilate the bathymetry of the lake, a three-dimensional hydrodynamic model based on FVCOM has been developed for the lake during the period of thermal stratification (JuneAugust of 2014). The hydrodynamic equations were solved by the numerical model on an unstructured grid and details are given in Table 2-1. 30 The surface water temperature was collected using an Onset HOBO Pro v2 sensor with an accuracy of 0.2 oC. A linearly varying startup water temperature was used with a value of 12 oC at the water surface and 4 oC at the depth of 10 m. The hydrodynamic model was tested using observed current data measured using a Teledyne - RDI Sentinel-V ADCP (1000 kHz frequency with a bin size of 0.3 m) deployed in the nearshore waters of the lake in approximately 10 m of water (Figure 2-7b). Finally, the bathymetry of the lake interpolated to grid nodes using the manifold method was assimilated into the model. 2.2.2 Reconstruction of meteorological data in Gull Lake 2.2.2.1 Land-based weather station network Hourly meteorological observations, including wind speed and direction, air temperature, air pressure, solar radiation, and relative humidity are needed for calculation of wind and heat flux fields. The meteorological observations were obtained from NCDC, Weather Underground (https://www.wunderground.com), and the Kellogg Biological Station Long-Term Ecological Research (KBS LTER, http://lter.kbs.msu.edu) stations, a total of 22 locations surrounding Gull Lake from May to August of 2014 and 2015 (Figure 2-8). Instead of a constant air pressure, hourly air pressure data recorded by the KBS LTER stations were used to improve the performance of the model. This also helped in the calculation of water density in FVCOM based on a polynomial expression [Jackett and Mcdougall, 1995] that takes pressure into account. Long wave solar radiation was estimated using air temperature and cloud cover [Parkinson and Washington, 1979]. 31 Figure 2-8 Selected weather stations surrounding Gull Lake. Since all weather stations are located within a 42 km radius of Gull Lake, a circular grid with a radius of 42 km was created for the reconstruction of meteorological fields (Figure 2-9). After applying overland-overlake adjustments, all observations were interpolated over the circular grid using a smoothed natural neighbor method. In order to have smooth and consistent wind and heat flux fields, a spatial moving averaged filter with a radius of 6 km was applied to the fields. This radius provided the best simulated results between the ranges of 0 to 30 km. Magnitudes of the smoothed fields were then adjusted based on the observations at the KBS LTER station to compensate for the reduction in magnitudes caused by the smoothing filter. Finally, the results were interpolated to the FVCOM’s computational grid using the natural neighbor interpolation method. 32 Figure 2-9 Circular grid with a radius of 42 km used for the reconstruction of meteorological fields from the land-based weather station network. 2.2.2.2 Mesoscale weather prediction model The Weather Research and Forecasting model (WRF3.7.1, http://www.wrf-model.org) with a lake physics component was used in this study to simulate meteorological forcing over Gull Lake. Three nested simulation domains with resolutions of 30 km, 10 km, and 2 km were defined and centered at (42.402778 oN, 85.41295 oW) on Gull Lake (Figure 2-10). 33 Figure 2-10 Three nested simulation domains with resolutions of 30 km, 10 km, and 2 km used for the WRF model. The 20-category MODIS-based land use data with a 30 s resolution of static data were interpolated to the model grids. USGS 24-category data were used as well, as an alternative set of land use, if a category from the MODIS-based data was not available. There are a number of meteorological reanalysis datasets which can be used as input to the WRF model. National Centers for Environmental Prediction / National Center for Atmospheric Research (NCEP/NCAR) reanalysis data provides four-times daily data with a spatial resolution of 206 km. The 32 km, three-hourly NCEP North American Regional Reanalysis (NARR) data can also be 34 used to generate initial and lateral boundary conditions. We selected Community Land Model Version 4 (CLM4, Oleson et al. [2010]) for the WRF’s land surface scheme in order to model the land - atmosphere interaction processes. The hourly wind field, air temperature, air pressure, shortwave and longwave solar radiations, and relative humidity were simulated from early April to September in years 2014 and 2015. 2.3 Results and Discussion 2.3.1 Manifold method True values of the analytical function at each of the randomly selected sampling locations were compared with the estimated values obtained by the manifold method as well as other standard interpolation methods. The performance statistics for this example are provided in Table 2-2. Definitions of metrics used in this study are provided in the Appendix. For all methods, the approximation of the F7 function improved by increasing the number of sampling points. In this particular example, the results show that the manifold method produced better overall performance compared to the other three methods considered. However, the best method in this example might perform differently on another test function or for a different sampling point distribution. Therefore, we examine the performance of the method for other datasets in the Lake Michigan and Gull Lake. Due to the sparse distribution of weather stations around Lake Michigan, it was not clear a priori how many neighboring stations would provide an adequate representation of the data. Since choosing a relatively few (e.g., three) neighboring stations in this situation would involve using information from stations that are far apart as neighbors, we used kernel regression to assign weights to each station depending on the distance from the point of interest. For each node of the 35 numerical grid of Lake Michigan, k number of nearest neighbors were selected and their assigned weights were projected to a low-dimensional subspace. Table 2-2 Cross-validation results for the analytical function based on different sampling points selected randomly. Sample size 30 60 90 R2 RMSE Fn PBIAS NASH APB (%) Manifold 0.667 0.778 0.710 14.420 0.416 0.652 Natural neighbor 0.582 0.876 0.799 -8.866 0.259 0.713 Nearest neighbor 0.619 0.882 0.804 -14.689 0.250 0.669 IDW 0.577 0.870 0.793 35.847 0.270 0.727 Manifold 0.846 0.579 0.531 -33.924 0.703 0.472 Natural neighbor 0.816 0.615 0.564 16.912 0.664 0.466 Nearest neighbor 0.779 0.720 0.660 -34.524 0.540 0.512 IDW 0.832 0.603 0.553 -44.292 0.677 0.469 Manifold 0.891 0.502 0.432 -14.103 0.791 0.400 Natural neighbor 0.874 0.539 0.464 -10.666 0.759 0.344 Nearest neighbor 0.867 0.571 0.491 -28.777 0.730 0.446 IDW 0.859 0.567 0.487 -5.974 0.735 0.416 Method The free parameters in the method are c (time scale), σ (the parameter used in kernel regression), and k. The standard deviation of weather station distances from the point of interest was used for the parameter  in kernel regression. Performance of the manifold method as measured by a comparison of simulated and observed currents in Lake Michigan is summarized in Table 2-3 relative to the other standard methods considered. We note that the manifold method based on kernel weighting considering all stations produced the best overall performance as measured by the root mean squared error (RMSE) between the observed and simulated currents. The performance of the method without kernel regression and with only three neighboring stations 36 was comparable to the other methods, but slightly inferior to the natural and nearest neighbor methods. Table 2-3 RMSE values (m/s) of alongshore and cross-shore velocities in Lake Michigan for comparison of the manifold method with other standard methods used in limnology and oceanography. Loaction: M Location: BB Location: S Method RMSEu RMSv RMSEu RMSEv RMSEu RMSEv O-kriging 0.0385 0.0290 0.0590 0.0349 0.0540 0.0152 Nearest Neighbor 0.0363 0.0286 0.0580 0.0348 0.0545 0.0152 Natural Neighbor 0.0366 0.0275 0.0553 0.0334 0.0515 0.0158 Manifold (3 NBR) 0.0383 0.0276 0.0594 0.0346 0.0568 0.0158 Manifold+Kernel (3 NBR) 0.0371 0.0268 0.0576 0.0341 0.0559 0.0158 Manifold+Kernel (all NBR) 0.0304 0.0265 0.0531 0.0312 0.0568 0.0154 IDW (all NBR) 0.0328 0.0267 0.0535 0.0316 0.0498 0.0155 Figure 2-11 shows the RMSE and R2 values for different numbers of nearest neighbors at different ADCP locations. Having all stations to create the tangent space for the manifold method resulted in a better representation of wind fields, and improved the results of the hydrodynamic model (Figure 2-12). Cross-validation was also used to compare the performance of the manifold method with other standard methods for the same Lake Michigan datasets. The performance metrics are summarized in Table 2-4. In this cross-validation method, one weather station was withheld to be used later for validating the manifold method, and all other stations surrounding the lake were used for the manifold training set. This process was repeated so that each weather station was given a chance to be part of this validation process. Based on these results the proposed manifold method with three nearest neighbors gave better results compared to other standard methods. However, the performance of the hydrodynamic model based on these methods was 37 relatively inferior compared to the performance of the model when the manifold method used all neighboring points. Figure 2-11 Performance of the manifold method evaluated using observed and simulated currents at different stations in Lake Michigan. Different number of nearest neighbors were used to reconstruct the wind field using the manifold method with kernel regression. Table 2-4 Cross-validation results for wind field over Lake Michigan. R2u R2v RMSEu RMSEv Computational time (s) O-kriging 0.441 0.572 3.497 3.853 92463.8 Nearest Neighbor 0.666 0.743 2.792 3.044 18.6 Natural Neighbor 0.693 0.794 2.558 2.750 183.6 Manifold (3 NBR) 0.690 0.801 2.433 2.595 28.1 Manifold+Kernel (3 NBR) 0.710 0.806 2.392 2.566 55.1 Manifold+Kernel (all NBR) 0.547 0.681 2.884 3.129 77.3 IDW (3 NBR) 0.724 0.822 2.278 2.458 69.3 Method All different versions of the manifold methods had reasonable computational efficiency. The computational time for the O-kriging was high due to the time needed for finding the best 38 variogram at each time step. Computations were performed using MATLAB, on an Intel Core i74771 3.5 GHz platform. As the last case study, wind and heat flux fields of the Gull Lake were interpolated over the numerical mesh of the lake using the natural neighbor method. Then, the bathymetry of the lake was interpolated over the mesh using the same natural neighbor method to develop the initial version of the lake hydrodynamic model. The raw bathymetry data, which has some regions of steep bathymetry change, created artificial currents in the model due to an error in the pressure gradient force introduced by the sigma-coordinate system of FVCOM [Mellor et al., 1998]. Therefore, the interpolated bathymetry was smoothed with a radius of 100 m in order to reduce the errors. Figure 2-12 Comparison of simulated and observed vertically averaged currents at the location M in Lake Michigan. (a) Alongshore velocity (b) Cross-shore velocity The developed model was used to assimilate the bathymetry of the lake based on the manifold method. First, the bathymetry data were reconstructed from the tangent space of the manifold with 39 three nearest neighbors and smoothed with the same method described above. Then, the hydrodynamic model was run with the reconstructed bathymetry. The comparisons of the vertically-averaged velocity profiles at the ADCP location using natural neighbor method, IDW with three nearest neighbors, and manifold method are presented in Figure 2-13. The best value of σ used in kernel regression was equal to the standard deviation of distances of observational points where water depth values are available within a search radius of 50 m from the point of interest. Figure 2-13 Comparison of simulated (black lines) and observed (red lines) vertically averaged currents at the ADCP location in Gull Lake. (a) Eastward velocity and (b) Northward velocity When the number of samples within this radius was smaller than 100, σ value was calculated based on locations of 100 nearest samples. This method is more accurate when enough samples are available around an estimated point, unless selecting 100 samples itself does a reasonable job. RMSE values (m/s) of eastward and northward velocities in Gull Lake for comparison of the manifold method with other standard methods used in limnology and oceanography are presented in Table 2-5. 40 Table 2-5 RMSE values (m/s) of eastward and northward velocities in Gull Lake for comparison of the manifold method with other standard methods used in limnology and oceanography. Method RMSE u RMSE v Natural Neighbor 0.0090 0.0205 Manifold+Kernel (3 NBR) 0.0098 0.0200 IDW (3 NBR) 0.0095 0.0204 The statistics of cross-validation for all (=71) measured longitudinal and transverse transects are shown in Table 2-6. The cross-validation was performed by omitting one transect at each step and calculating the bathymetry for that transect from the rest of the observation data and repeating the process for all other transects. Table 2-6 Cross-validation results for Gull Lake bathymetry. R2 RMSE (m) Fn NASH PBIAS Manifold 0.890 2.011 0.222 0.678 -14.016 Natural Neighbor 0.925 1.288 0.170 0.468 -13.301 Nearest Neighbor 0.888 2.039 0.230 0.670 -17.132 IDW (3 NBR) 0.839 3.282 0.6065 0.540 -15.918 Method 2.3.2 Observed and predicted meteorological data in Gull Lake Two types of reanalysis data (NCEP/NCAR and NARR) were tested for use as input data of the WRF model. The WRF simulated weather data were compared with the observed data at the KBS LTER station to evaluate the performance of the weather forecast model. Comparisons of observed meteorological data at KBS LTER station with simulated results based on NCEP/NCAR are presented in Figure 2-14. Figure 2-15 also shows scatter plots of WRF-simulated results based on NARR data versus observed meteorological forcing at KBS LTER. The simulated results, particularly wind speed and air temperature, indicated that NARR data would be a better choice 41 for the input of the hydrodynamic model. Thus, NARR data will be utilized in Chapter 3 to simulate the meteorological forcing for the coupled WRF-lake model. The results showed promising performance in simulation of the weather data at the KBS LTER station. Comparisons of averaged simulated forcing and meteorological forcing over Gull Lake during April to midSeptember of 2014 and 2015 are presented in Figure 2-16 and 2-17, respectively. Figure 2-14 Comparisons of observed meteorological data at KBS LTER station with simulated results based on NCEP/NCAR data during April to mid-September of 2014 (a) air pressure (kPa), (b) air temperature ( C ), (c) downward shortwave radiation (W/m2), (d) downward longwave radiation (W/m2), (e) relative humidity (%), (f) wind speed at 10-meter height (m/s), (g) eastward wind velocity (m/s), (h) northward wind velocity (m/s). Since the raw wind fields reconstructed using weather station data alone were in good agreement with the overlake WRF model results, no overlake-overland adjustment [Schwab and Morton 1984] was applied to the meteorological forcing. An empirical overland-overlake adjustment has usually been applied to wind speeds recorded by land-based weather stations to determine overlake wind speed in the Great Lakes [Schwab and Beletsky, 1998; Thupaki et al., 2013; Nguyen et al., 2014; Safaie et al., 2016]. However, this study shows that the overlandoverlake adjustment can generate high values of overlake wind speeds for a small inland lake, 42 which has a short fetch length compared to large lakes. On the other hand, the primary results of the coupled WRF-lake model show that current speeds were overestimated due to the relatively high simulated wind speed. Figure 2-15 Scatter plots of WRF-simulated results versus observed meteorological forcing at KBS LTER during April to mid-September of 2014 (black dot) and 2015 (red dot): (a) air pressure (kPa), (b) air temperature ( C ), (c) downward shortwave radiation (W/m2), (d) downward longwave radiation (W/m2), (e) relative humidity(%), (f) wind speed at 10-meter height (m/s), (g) eastward wind velocity (m/s), (h) northward wind velocity (m/s). A comparison of the simulated and observed wind speeds for years 2014 and 2015 at the KBS LTER station is shown in Figure 2-18. These comparisons indicate that for both 2014 and 2015, the simulated wind speeds are about 30 percent higher than observations. WRF is known to over predict the wind speed depends on the topographic complexity, drag parameterization, and its vertical and horizontal resolutions [Jiménez and Dudhia, 2012; Brunner et al., 2015; Staffell and Pfenninger, 2016]. Therefore, an adjustment factor of 0.7 should be applied into the WRFsimulated wind speeds to be used as inputs of the Gull lake model. It is also worth mentioning that initially the shortwave solar radiation was calculated using theoretical estimates of clear-sky solar 43 radiation [Annear and Wells, 2007], and adjusted by an empirical relation between the clear-sky value and the measured cloud cover [Bunker, 1976] to assess the WRF results. Since the nearest weather station (a NCDC station) with the cloud cover data was 23 km far from the lake, the calculated shortwave solar radiation becomes 30% smaller compared with WRF-simulated results. However, using the observed solar radiation data that were obtained from the KBS LTER station resulted in a better agreement with prediction of WRF. Figure 2-16 Comparisons of averaged simulated forcing (WRF) and observed meteorological forcing (Stations) over Gull Lake during April to mid-September of 2014. 44 Figure 2-17 Comparisons of averaged simulated forcing and meteorological forcing over Gull Lake during April to mid-September of 2015. Figure 2-18 Comparisons of the simulated and observed wind speed in 2014 and 2015 at the KBS LTER station. 45 2.4 Conclusions We presented a novel manifold method of reconstructing spatio-temporal data, which could be used for assimilating geophysical and meteorological data integrated land surface subsurface, and lake models. All case studies illustrate the superior performance of the presented manifold algorithm over standard methods in terms of accuracy and computational efficiency. The hydrodynamic model of Lake Michigan based on the manifold method of reconstructing wind fields produced better performance relative to the other methods. The best results were obtained using kernel regression applied to all weather stations (neighbors). However, the cross-validation results show that the results of the three nearest neighbors were better than the other methods. We can see that the manifold method performs better than the IDW method at two of the stations (M and BB) but not at the nearshore point S. We believe that the reason for this has to do with the fact that in the nearshore region there are a number of additional processes (waves, wave-current interactions etc) which are not simulated in our model. In other words, in that region the flow is not directly driven by the wind fields but rather indirectly through exchange between offshore and nearshore water and wave propagation. Therefore model performance in that region cannot be directly related to the wind field. At the other two offshore stations M and BB, the flow is predominantly wind-driven and an improvement in the simulated hydrodynamic fields can be seen. This also brings us to two relevant points: (1) Details of the manifold method such as the tangent space estimation, the distance metric that defines spatiotemporal proximity and other details can all be further improved to improve the performance of the manifold method, but these topics are beyond the scope of the present study. (2) We do not claim that the manifold method provides superior performance on all datasets and for all performance metrics, but from the examples considered here it appears that the manifold method may offer an attractive method that 46 is comparable or superior to other standard methods. Clearly more research is needed to understand the relative strengths and weaknesses of different manifold-based approaches compared to standard methods. The Gull Lake model results indicated that the proposed method has the ability to reconstruct geophysical data at unsampled locations. The results highlight that evaluating the performance of interpolation methods using the cross-validation method without considering the data structure and the purpose of interpolating data can lead to misleading conclusions about the relative performance of the methods considered. The WRF model was found to be predictive of meteorological data. Given the uncertainty involved due to the sparse distribution of weather stations, lack of quality assurance of raw weather data, and the overland-overlake effect, the weather prediction model could be utilized to assess the reconstructed meteorological forcing. The fact that meteorological forcing based on the outputs of a mesoscale weather prediction model (WRF) could provide results comparable to the forcing based on a network of weather stations is encouraging. However, further examination is needed to assess the accuracy of the WRF in representing the spatio-temporal fields of interest. Indeed, we can use the same approach presented here for evaluating the performance of the manifold method. Therefore, in the next chapter, a coupled WRF-lake model will be used to further assess the accuracy of meteorological forcing reconstructed from land-based weather stations and vice versa based on their ability to describe circulation and thermal structure of Gull Lake. 47 CHAPTER 3 3 Evaluating the Role of Groundwater in Circulation and Thermal Structure 3.1 Introduction Lake ecosystems readily respond to the influences of climate through a number of coupled surface and subsurface processes within their watersheds and therefore serve as sentinels of climate change [Adrian et al., 2009; Schindler, 2009]. A recent global synthesis of in situ and remotelysensed lake data indicated that lake surface temperatures increased quickly between 1985 and 2009 at a mean global rate of 0.34 C per decade. Similar warming trends were noted in streams, rivers and shallow groundwater as well [Kaushal et al., 2010; Menberg et al., 2014], although the rates of warming can be expected to be different for surface and groundwater systems. Due to the higher heat capacity of soils and the buffering effects of vegetation, groundwater temperatures remain relatively constant throughout the year partly offsetting the effects of surface warming in groundwater-fed lake ecosystems. While the effects of climate change on the Laurentian Great Lakes are well-documented (e.g., growing annual average temperatures, shorter winters, decreasing lake ice covers and ice-albedo feedback) and continue to receive significant attention (e.g., Austin and Colman [2007]; Nguyen et al. [2014]), there is growing evidence to indicate that 48 smaller inland lakes may respond differently to climate change than the Great Lakes of the world [Winslow et al., 2015]. Temperature is a key driver that affects the structure and function of ecosystems including biological productivity. Therefore, important changes related to nutrient cycling, eutrophication and hypoxia can be linked to changes in lake circulation and thermal structure. Most lake models, however, do not explicitly consider the effects of groundwater and quantifying these processes for small lakes presents unique challenges. For example, choices associated with the representation of meteorological forcing, lake morphometry, and turbulent mixing within the water column can give rise to model uncertainties in temperatures comparable to the global mean warming rate noted above. Therefore, there is a need to understand and quantify the role of groundwater in circulation and thermal structure within lake ecosystems beyond simple water budgets. In particular, there is a need to systematically evaluate the impact of several modeling choices on model outcomes in order to identify the best choices relative to the spatial and temporal scales of interest. In inland shallow lakes, heat exchange between water and lake sediment was taken into consideration for vertical thermal diffusivity analyses [Hondzo et al., 1991; Hondzo and Stefan, 1993]. Although, the role of groundwater in circulation and thermal structure of lakes are often assumed to be negligible, a shallow groundwater-fed lake may have significant bottom heating during winter and bottom cooling in summer [Kettle et al., 2012] Relatively small lakes such as Gull Lake, the focus of this study, have shorter fetch lengths compared to large lakes and wind speed can be significantly decreased due to sheltering effects [Hondzo and Stefan, 1993; Markfort et al., 2010]. Therefore, wind- and wave-driven turbulent mixing are weaker in small lakes, and consequently, influences of water clarity on mixed-layer depth of small lakes are more pronounced than in large lakes [Heiskanen et al., 2015]. Physical, 49 chemical, and biological processes at the air-water interface, mixed layer, and thermocline are directly influenced by the distribution of solar radiation in the water column [Simpson and Dickey, 1981b]. For example, phytoplankton competition for nutrients and light in a stratified water column is controlled by the ratio of light attenuation to their nutrient consumption [Yoshiyama et al., 2009]. A minor change in parameter values such as the mixed layer depth or light attenuation length may drastically change the vertical distribution of phytoplankton [Mellard et al., 2011]. Accurate mechanistic models capturing physical aspects of these processes hold the key to the understanding and predicting biophysical processes. Despite the availability of high-quality data based on decades of water quality monitoring and biological research in Gull Lake, a numerical hydrodynamic model of the lake has not been developed to date. One of the objectives of this chapter is to fill this gap as a first step towards building a coupled physical-chemical-biological modeling system. The aim of this study was to evaluate the role of groundwater in circulation and thermal structure within Gull Lake, which is typical of many other groundwater-fed inland lakes in Michigan. Groundwater effects on lakes can be studied at different spatial and temporal scales. The focus of the present work is on examining lake hydrodynamics and thermal structure during the summer stratified period using a “lake modeling perspective” in which the effects of groundwater are introduced via boundary conditions without explicitly coupling with groundwater models. We seek to address the following questions as part of this research: (a) How will summer stratification and circulation change if groundwater contribution is ignored? This question is important as groundwater levels are declining in many parts of the world reducing or eliminating groundwater contributions to lakes. (b) Can mesoscale weather prediction model outputs provide sufficiently accurate forcing fields to run hydrodynamic models for relatively small lakes? If so, 50 how do results compare with those obtained using observations from a network of weather stations? This question is important since many small lakes throughout the world do not have a network of meteorological stations around them. (c) To what extent do alternative turbulence models such as the two-equation k-ε model [Rodi, 1987] improve descriptions of vertical mixing and thermal structure in small lakes such as the Gull Lake? The Mellor-Yamada 2.5 level turbulence model (MY2.5, Mellor and Yamada [1982]) is often used to describe vertical mixing in lake and ocean circulation models. The performance of the k-ε turbulence model for small, groundwater-fed lakes will be evaluated to identify the best model. The chapter is structured as follows. Following a description of the study site, we describe data collected in the lake to evaluate the performance of the numerical models. After a description of the hydrodynamic model, we describe how the numerical mesh was refined using a bathymetric map of the lake for a more accurate simulation of thermal structure. Surface heat fluxes are usually the primary sources driving heat transfer in the system; therefore, we describe an approach to assess the accuracy of our station-based meteorological forcing by coupling a mesoscale weather prediction model with our lake model. Model descriptions of vertical turbulent mixing, a key process for transferring heat downward and for the onset of stratification, may not be accurate enough under conditions of strong stratification [Li et al., 2005]. Therefore, we compare the performance of two turbulence models. Changes in internal heating of the water column associated with fluctuations in water clarity and photic zone depth are known to influence stratification and vertical mixing [Chen, 2003a]. Therefore, we describe the use of in situ observations of photosynthetically available radiation (PAR) to enhance the performance of the shortwave penetration model in the hydrodynamic model. After these improvements, we describe how 51 groundwater effects are simulated in the lake model. Finally, an assessment of simulation results is presented along with a discussion and conclusions. 3.2 Materials and methods 3.2.1 Observational data Field observations were made during the summer months of 2014 and 2015 to test the hydrodynamic model. Bathymetry data were collected along longitudinal and transverse transects of the lake using a SonTek RiverSurveyor M9 system, as described in Chapter 2. Figure 3-1 is a map of Gull Lake with locations of these in situ measurements marked. In the summer of 2014, we deployed, for the first time in the lake, two upward-looking ADCPs manufactured by Teledyne RD Instruments (at locations marked M14 and S14 in Figure 3-1 and obtained high-resolution current and temperature data. A thermistor chain using Onset HOBO Pro v2 temperature sensors (Onset Computers Inc, Cape Cod, Massachusetts, USA) with an accuracy of 0.2 oC was deployed near the ADCPs (at location marked TC in Figure 3-1) in 15 m of water. This thermistor chain had sensors between 2-13 m depth of waters with a one-meter interval (Figure 3-2). We also had a surface buoy near TC location to record surface water temperature. A Hydrolab multi-parameter sonde (OTT Hydromet, Kempten, Germany) was used to measure the vertical temperature profile at another location (marked HL in Figure 3-1) in 32 m of water. In 2015, two other ADCPs were deployed at locations SS15 and SM15 for 91 days from early June to early September. A SCAMP (Self-Contained Autonomous Micro-Profiler, http://pme.com) was used to collect salinity and photosynthetic active radiation (PAR) data at the center of Gull Lake (HL location) during the summers of 2014 and 2015. Further details of instruments deployed in Gull Lake are provided in Table 3-1. 52 Figure 3-1 Map of Gull Lake showing locations of the in situ measurements. Table 3-1 Details of instruments deployed in Gull Lake. ID Instrument Year Depth Deployment Location Latitude Longitude M14 600 kHz Monitor ADCP 2014 11.6 42.4047 -85.403 S14 1000 kHz Sentinel V20 ADCP 2014 9.45 42.3963 -85.416 42.402 -85.407 SD15 1000 kHz Sentinel V20 ADCP 2015 19 SS15 1000 kHz Sentinel V20 ADCP 2015 10.4 42.4051 -85.404 TC Thermistor Chain 2014-2015 12.96 42.4054 -85.404 HL Hydrolab and SCAMP 2014-2015 32 42.3959 -85.407 53 Figure 3-2 Schematic of the thermistor chain. 3.2.2 Numerical hydrodynamic model The mechanistic model of Gull Lake has been developed based on a three-dimensional, unstructured grid, Finite-Volume Community Ocean Model (FVCOM; Chen et al. [2003b, 2006]). FVCOM has been successfully applied to all Great Lakes of North America [Bai et al., 2013], including Lake Michigan [Luo et al., 2012; Rowe et al., 2015; Safaie et al., 2016, 2017a], Lake Huron [Nguyen et al., 2014], Lake Superior [Xue et al., 2015], Lake Erie [Jiang et al., 2015; Niu et al., 2015], Lake Ontario [Shore, 2009; Wilson et al., 2013] and large rivers such as the St. Clair 54 River [Anderson and Phanikumar, 2011]. The Reynolds-averaged Navier–Stokes equations of momentum (Eq. 1-3), continuity (Eq. 4), temperature (Eq. 5), and salinity (Eq. 6) are as follows: u t u v t u w t u u x u x v v x v w x v v y u y w v y w w y w w z 0 T t u T x v T y S t u S x v S y where u , v , and u z fv v z fu w z 1 0 1 0 1 P z (P ) x z (P ) y g z z (Km u ) z Fu (1) (Km u ) z Fv (2) (Km w ) Fw z (3) (4) w w T z S z z z (Kh (Kh T ) FT z S ) FS z (5) (6) w are mean velocity components in x , y , and z directions, respectively. T and S are time-averaged temperature and salinity. denotes the mean density and 0 is the reference density. P is the mean pressure. f is the Coriolis parameter and g is the gravitational acceleration. K m and K h are vertical eddy viscosity and thermal vertical eddy diffusion coefficients, respectively. Fu and Fv are the horizontal momentum; Fw , FT and FS are vertical momentum, thermal, and salt diffusion terms. By a scaling argument for vertical velocity, the vertical momentum equation (Eq. 3) is reduced to the following hydrostatic equation: 55 1 P z 0 g Fu , Fv , FT and FS are modeled using the following forms: Fu Fv FT FS x y x x (7) (2Am u ) x (2Am v ) y (Ah T ) x (Ah S ) x y x y y (Am ( (Am ( u y v x v )) x (8) u )) y (9) (Ah T ) y (10) (Ah S ) y (11) where Am and Ah are horizontal diffusion coefficients for momentum and scalars, respectively. The governing equations of FVCOM as the default are closed with the modified Mellor - Yamada 2.5 level [Mellor and Yamada, 1982] and Smagorinsky turbulent closure schemes for vertical and horizontal mixing, respectively. A two-equation turbulence model [Rodi, 1987] was utilized as an alternative turbulence closure scheme for vertical mixing in this work. For this purpose, the General Ocean Turbulence Model (GOTM4.0, http://www.gotm.net), originally developed by Burchard et al. [1998] was coupled to FVCOM. Details of the turbulence closure models and parameter values are described in Section 2.3. The three-dimensional hydrodynamic equations of the lake were solved using a mode splitting method [Simons, 1974; Madala and Piacseki, 1977]. In this method, the vertically integrated equations (external mode) and the vertical structure equations (internal mode) are solved separately with different time steps. Since the internal waves or the mean currents travel much 56 more slowly than the external mode, a larger time step can be used for the internal mode to reduce the computational time. Initial conditions for water temperature in the lake used the measured temperature profiles obtained using the Hydrolab sonde. For salinity we used the SCAMP data. The mean density, which is a function of water temperature, salinity, and pressure [Jackett and Mcdougall, 1995], was recalculated in pressure coordinates every 30 minutes. This recalculation of the mean density helps by producing more stable stratification. Average depth of the lake is approximately 12.5 m with a maximum water depth of 33 m. While typical lake-bed slopes in Gull Lake are less than 0.2, the bottom slopes at some points, especially near the deployment sites, are as steep as 0.55. To follow the bottom topography, FVCOM, which is a terrain-following model, uses a vertical σ-coordinate transformation. Although the advantages of the σ-coordinate system are well-known, this coordinate system introduces pressure gradient force errors over steep topography [Mellor et al., 1998]. Previous research showed that bathymetry smoothing, to some extent, would be required to achieve stability and accuracy in simulations with an adequate representation of topography [Barnier et al., 1998; Haidvogel et al., 2000; Martinho and Batteen, 2006; Sikirić et al., 2009]. In Chapter 2, we used a moving average filter with a radius of 100 m to smooth the bathymetry and to reduce artificial noise in simulated currents. However, detailed bathymetry data with fine horizontal and vertical grid resolutions are needed to simulate hydrodynamics and thermal structure accurately. Similar to bottom steepness, vertical and horizontal resolutions, and the stratification strength have impacts on the pressure gradient errors [Haidvogel et al., 2000], and because Gull Lake exhibits strong stratification during summer months, adequate mesh resolution is a key to the success of numerical modeling. One way to avoid pressure gradient errors in σ-coordinates is to generate a high-quality unstructured computational mesh [Gorman et al., 2006]. We used the Surface Water Model 57 System software (SMS 11.1, http://www.aquaveo.com) developed by Zundel and Jones [1996], to generate a triangular horizontal mesh. We used the following bathymetry-based algorithm to resolve topographic features and to reduce the pressure gradient errors. First, we created contour lines with one meter elevation interval from the collected bathymetry. Then, we selected contour lines that follow the key topographic features of the lake, and treated them as constraint lines, along which grid nodes must be placed. The unstructured mesh was generated to meet the quality criteria for maximum slope. Elements with slopes greater than 0.2 were refined. A maximum slope of 0.1 is recommended for FVCOM [Chen et al., 2006], although it is preferable to use larger values (such as 0.3) to retain the accuracy of the raw bathymetry data [Foreman et al., 2009]. The computational mesh was also refined where the slope parameter, recommended by Mellor et al. [1994], exceeded the limit of 0.2. The slope parameter for each element is defined as: r h (12) 2h where h is the maximum depth difference between adjacent grid points, and h is the average of the depths for the two adjacent grid cells. We limited the minimum grid resolution to 5 m to avoid high computational cost and numerical instability issues. Therefore, in regions where either maximum slope or slope parameter conditions needed a finer resolution, the bathymetry was smoothed using the method described in Mellor et al. [1998]. A numerical mesh with 20, 604 nodes and 40, 260 triangular elements in the horizontal with a resolution range of 5-75 m was created following the bathymetry-based refinement algorithm (Figure 3-3). In the vertical, a sigma coordinate system with 30 sigma layers was used. 58 Figure 3-3. The unstructured computational mesh created based on the bathymetry-based refinement algorithm The hydrodynamic model of Gull Lake based on the meteorological forcing for summer 2014 was run for three different cases. To systematically quantify the improvements in model performance due to the choices in reconstructing meteorological forcing fields, turbulence models and inclusion of groundwater, precipitation and evaporation processes, several cases (described in Table 3-2) were considered. First the model was run with the k-ε turbulence model and without the groundwater module (Model 1). Then for the second case, the groundwater module was taken into account in order to examine the effect of groundwater on the physical behavior of the lake (Model 2). This case also was run with the MY2.5 turbulence model (Model 3). Since our primary interest is in evaluating the role of groundwater in circulation and thermal structure of Gull Lake, the coupled WRF-lake model was run by including the contribution from groundwater and with 59 Table 3-2. Description of different models evaluated Model description 2014 Year Meteorological Turbulence Forcing Model Model1 Weather Stations k-ε --- Model2 Weather Stations k-ε GW Together with model 3, simulation provides a comparison of two popular turbulence models for their ability to describe vertical mixing Model3 Weather Stations MY2.5 GW Together with model 2, simulation provides a comparison of two popular turbulence models for their ability to describe vertical mixing k-ε GW Together with model 2 which is based on weather station data, simulation allows a comparison of results based on two different meteorological forcing fields (weather station data versus WRF model outputs) 2015 Model4 WRF Model Additional Processes* Simulation shows the effect of ignoring groundwater. Model1 Weather Stations k-ε --- Same as model 1 but for year 2015 Model2 Weather Stations k-ε GW Same as model 2 but for year 2015 k-ε GW Same as model 4 but for year 2015 Model4 WRF Model * Remark Model name GW = Groundwater 60 the k-ε turbulence model (Model 4). At the end, the developed models were also tested using data obtained during the summer of 2015. Details of the reconstruction of meteorological forcing fields, model of shortwave radiation penetration and the groundwater-precipitation-evaporation module are explained in the following sections. 3.2.3 Turbulent closure models 3.2.3.1 Mellor-Yamada 2.5 level turbulence model The governing equations of FVCOM, including momentum, continuity, temperature, salinity, and density equations, as the default are closed with the modified Mellor and Yamada level 2.5 (MY-2.5; Mellor and Yamada [1982]) and Smagorinsky turbulent closure schemes for vertical and horizontal mixing, respectively. The Smagorinsky horizontal diffusion for momentum is given as: Am u x 0.5C 2 0.5 v x u y 2 v y 2 (13) where C is a constant parameter and Ω is the area of the individual momentum control element. Then, the horizontal diffusion coefficients for scalars can be calculated as Am / Ah Pr , where Pr is the turbulent Prandtl number. MY-2.5 is a two equation model that solves transport equations for variables q 2 (twice of turbulent kinetic energy) and q 2l , where l is the turbulence length scale. The set of q ql equations in FVCOM is defined as [Chen et al., 2003b]: q2 t u q2 x v q2 y w q2 z 2(Ps Pb ) z 61 (Kq q2 ) Fq z (14) q 2l t q 2l x u q 2l y v w q 2l z lE1 (Ps W ) E1 Pb z (Kq q 2l ) z Fl (15) w are components of the mean velocity in x , y , and z directions. Fq and where u , v , and Fl denote the horizontal diffusion of turbulent kinetic energy and turbulence macroscale, respectively. Kq is the vertical eddy diffusion coefficient and q 3 / B1l is the dissipation rate of turbulent kinetic energy, where B1 is a model constant. W 1 E2l 2 / ( L)2 is the wall proximity function where L 1 ( z) (H 1 z) 1 , is free surface elevation, H is mean water depth, and is the von Kármán constant, equal to 0.4. The constant parameters of the q ql model are (1.8, 1.33) . Ps (E1 , E 2 ) Ps Pb uw g u z w 0 vw Kh ( and Pb are shear and buoyancy production terms, respectively: v z Km [( g 0 z ) u 2 ) z ( v 2 ) ] z KmM 2 (16) Kh N 2 (17) where u , v , and w are turbulent velocity components and N denotes the Brunt-Väisälä frequency, M is the shear frequency, is the mean water density, 0 is the reference density, and g is the acceleration due to gravity. To close the system Eqs. (14) and (15), the vertical eddy viscosity coefficient ( K m ), the thermal vertical eddy diffusion ( K h ) and Kq are modeled as: Km lqSm , Kh lqSh , Kq lqSq (18) 62 where Sm , S h , and Sq are the stability functions originally proposed by Mellor and Yamada [ 1982] and simplified by Galperin et al. [1988] as the following [Allen et al., 1995]: Sh A2 Sm A1 Sq 1 6AB 1 1 1 3A2 (6A1 (1 3C1 1 (19) B2 )GH 1 6AB 9(2A1 1 1 ) 1 9AAG 1 2 H A2 )S HGH (20) Sm where GH (21) l 2N 2 / q 2 restricted to the maximum value of 0.023 for unstable stratification conditions, and the minimum value of -0.28 for conditions of stable stratification [Galperin et al., 1988]. The values of the constant parameters A1 , B1 , A2 , B2 , andC 1 are 0.92, 16.6, 0.74, 10.1, and 0.08 [Mellor and Yamada, 1982], respectively. The surface and bottom boundary conditions for the turbulent kinetic flux are given as: 2 ql 0, q ql 0, q 2 2 2 2 3 1 (x , y , t ) (22) 2 3 1 H (x , y , t ) (23) B u 2s at z B u 2b at z where u s and u b are surface and bottom friction velocities, respectively. 3.2.3.2 The k turbulence model The simplified form of the transport equations for the turbulent kinetic energy ( k ) and the rate of dissipation ( ) are given as following: 63 k t t k x u u x where k v v k y k z z x x z t ( k ( k ) P z t z ) k B (24) (c 1P c 3B c 2 ) are constant turbulent Schmidt numbers of k and , respectively. and (25) t denotes the eddy viscosity; P and B are the shear and buoyancy production terms, respectively, with the aforementioned definitions in the MY2.5 model. The turbulent diffusivities of momentum and heat can be expressed as: t c k 1/2l , t c k 1/2l (26) where c and c are stability functions computed according to the model of Schumann and Gerz [ 1995]: c c0 c0 , c (27) Prt where c is a model constant, and Prt is the turbulent Prandtl number with the following empirical 0 relation: Prt Prt0 exp( Ri Ri ) 0 Prt Ri Ri (28) 0 where Ri denotes the gradient Richardson number, Prt is the turbulent Prandtl number for neutral flows ( Ri 0 0 ), and Ri is the desired steady-state Richardson number. The values of Prt and Ri were set as 4 and 0.25, respectively. The constant parameters of the k 64 model are c0 0.5577 , 1.0 , k 1.3 , c 1.44 , 1 and c 2 1.92 [Rodi, 1987]. The parameter c 3 can be obtained as a function of a prescribed steady-state Richardson number, Rist [Umlauf and Burchard, 2003]: c 3 c 2 (c 1 c 2 ) c 1 c Rist (29) 3.2.4 Light attenuation lengths To accurately represent the thermal structure, surface temperatures and the depth of thermocline, it is important to correctly model how sunlight penetrates into water. In particular, absorption of downward irradiance in the stratified region has a strong influence on water temperature to the extent that different shortwave penetration models produce different stratification regimes and residual currents [Chen, 2003a]. In FVCOM, the depth distribution of the downward shortwave solar radiation flux, SW (x , y , z ,t ) , is calculated as: SW (x , y , z , t ) SW (x , y , , t ) R e za (1 R) e zb (30) where SW(x, y, z, t) is the downward shortwave radiation that penetrates into water as a function of x, y, water depth (z), and time (t). SW(x, y, ζ, t) represents the downward shortwave radiation at the water surface (z = ζ). R is an empirical constant, a is a short wavelength due to the near-surface absorption of red spectral components of solar radiation in the upper few meters, and b is a long wavelength related to the blue-green spectral components in a deeper water depth (a3 mg O2 L-1 and 8 oC 2.0.CO;2. Ambrose, R. B., T. A. Wool, and J. L. Martin (1993), The Water Quality Analysis Simulation Program, WASP5, Part A: Model Documentation, Environmental Research Laboratory, U.S. Environmental Protection Agency, Athens, Georgia. Anderson, M. P. (2005), Heat as a Ground Water Tracer, Ground Water, 43(6), 951–968, doi:10.1111/j.1745-6584.2005.00052.x. Anderson, E. J., and M. S. Phanikumar (2011), Surface storage dynamics in large rivers: Comparing three-dimensional particle transport, one-dimensional fractional derivative, and multirate transient storage models, Water Resour. Res., 47(9), 1–15, doi:10.1029/2010WR010228. Annear, R. L., and S. A. Wells (2007), A comparison of five models for estimating clear-sky solar radiation, Water Resour. Res., 43(10), 1–15, doi:10.1029/2006WR005055. APHA (1998), Standard methods for the examination of water and wastewater, 20th ed., American Public Health Association, Washington, DC. Apple, B. A., and H. W. Reeves (2007), Summary of Hydrogeologic Conditions by County for the State of Michigan, Open-File Report, U.S. Geological Survey, Reston, Virginia, United States. 135 Asem, A., A. Eimanifar, M. Djamali, P. De los Rios, and M. Wink (2014), Biodiversity of the Hypersaline Urmia Lake National Park (NW Iran), Diversity, 6(1), 102–132, doi:10.3390/d6010102. ASTM D888 (2012), Standard Test Methods for Dissolved Oxygen in Water, ASTM International, West Conshohocken, PA. Austin, J. (2013), Observations of near-inertial energy in Lake Superior, Limnol. Oceanogr., 58(2), 715–728, doi:10.4319/lo.2013.58.2.0715. Austin, J. A., and S. M. Colman (2007), Lake Superior summer water temperatures are increasing more rapidly than regional air temperatures: A positive ice-albedo feedback, Geophys. Res. Lett., 34(6), doi:10.1029/2006GL029021. Bai, X., J. Wang, D. J. Schwab, Y. Yang, L. Luo, G. A. Leshkevich, and S. Liu (2013), Modeling 1993–2008 climatology of seasonal general circulation and thermal structure in the Great Lakes using FVCOM, Ocean Model., 65, 40–63, doi:10.1016/j.ocemod.2013.02.003. Baines, S. B., K. E. Webster, T. K. Kratz, S. R. Carpenter, and J. J. Magnuson (2000), Synchronous behavior of temperature, calcium, and chlorophyll in lakes of northern Wisconsin, Ecology, 81(3), 815–825. Barnier, B., P. Marchesiello, A. P. De Miranda, J.-M. Molines, and M. Coulibaly (1998), A sigmacoordinate primitive equation model for studying the circulation in the South Atlantic. Part I: Model configuration with error estimates, Deep Sea Res. Part Oceanogr. Res. Pap., 45(4), 543–572. Barth, A., A. A. Azcárate, P. Joassin, B. Jean-Marie, and C. Troupin (2008), Introduction to Optimal Interpolation and Variational Analysis, SESAME Summer School, Varna, Bulgaria. Belkin, M., and P. Niyogi (2003), Laplacian eigenmaps for dimensionality, Speech Commun., 1(2– 3), 349–367. Biggs, R. et al. (2012), Toward Principles for Enhancing the Resilience of Ecosystem Services, Annu. Rev. Environ. Resour., 37(1), 421–448, doi:10.1146/annurev-environ-051211123836. Blumberg, A. F., and G. L. Mellor (1987), A description of a three-dimensional coastal ocean circulation model, Am. Geophys. Union, (edited by N. Heaps), 1–16. Boegman, L., M. R. Loewen, P. F. Hamblin, and D. A. Culver (2008), Vertical mixing and weak stratification over zebra mussel colonies in western Lake Erie, Limnol. Oceanogr., 53(3), 1093–1110. Brewer, M. K. (1991), A regional study of ground-water quality in Barry County, Michigan, M.S., Western Michigan University, United States -- Michigan. 136 Brooks, A. S., and J. C. Zastrow (2002), The Potential Influence of Climate Change on Offshore Primary Production in Lake Michigan, J. Gt. Lakes Res., 28(4), 597–607, doi:10.1016/S0380-1330(02)70608-4. Bruesewitz, D. A., J. L. Tank, and S. K. Hamilton (2009), Seasonal effects of zebra mussels on littoral nitrogen transformation rates in Gull Lake, Michigan, U.S.A., Freshw. Biol., 54(7), 1427–1443, doi:10.1111/j.1365-2427.2009.02195.x. Brunner, P., C. T. Simmons, and P. G. Cook (2009), Spatial and temporal aspects of the transition from connection to disconnection between rivers, lakes and groundwater, Journal of Hydrology, 376(1–2), 159–169, doi:10.1016/j.jhydrol.2009.07.023. Brunner, D. et al. (2015), Comparative analysis of meteorological performance of coupled chemistry-meteorology models in the context of AQMEII phase 2, Atmos. Environ., 115, 470–498, doi:10.1016/j.atmosenv.2014.12.032. Bunker, A. F. (1976), Computations of Surface Energy Flux and Annual Air–Sea Interaction Cycles of the North Atlantic Ocean, Mon. Weather Rev., 104(9), 1122–1140, doi:10.1175/1520-0493(1976)104<1122:COSEFA>2.0.CO;2. Burchard, H., O. Petersen, and T. P. Rippeth (1998), Comparing the performance of the MellorYamada and the κ-ε two-equation turbulence models, J. Geophys. Res. Oceans, 103(C5), 10543–10554, doi:10.1029/98JC00261. Burns, N. M., D. C. Rockwell, P. E. Bertram, D. M. Dolan, and J. J. H. Ciborowski (2005), Trends in Temperature, Secchi Depth, and Dissolved Oxygen Depletion Rates in the Central Basin of Lake Erie, 1983–2002, J. Gt. Lakes Res., 31, 35–49, doi:10.1016/S03801330(05)70303-8. Carin, L., R. G. Baraniuk, V. Cevher, D. Dunson, M. I. Jordan, G. Sapiro, and M. B. Wakin (2011), Learning Low-Dimensional Signal Models, IEEE Signal Process. Mag., 28(2), doi:10.1109/MSP.2010.939733. Carpenter, S. R., and M. G. Turner (2000), Hares and Tortoises: Interactions of Fast and Slow Variablesin Ecosystems, Ecosystems, 3(6), 495–497, doi:10.1007/s100210000043. Carpenter, S. R., E. H. Stanley, and M. J. Vander Zanden (2011), State of the World’s Freshwater Ecosystems: Physical, Chemical, and Biological Changes, Annu. Rev. Environ. Resour., 36(1), 75–99, doi:10.1146/annurev-environ-021810-094524. Chang, C.-H., L.-Y. Cai, T.-F. Lin, C.-L. Chung, L. van der Linden, and M. Burch (2015), Assessment of the Impacts of Climate Change on the Water Quality of a Small Deep Reservoir in a Humid-Subtropical Climatic Region, Water, 7(4), 1687–1711, doi:10.3390/w7041687. Chen, C., Beardsley, R.C., Franks, P.J.S., Van Keuren, J., (2003a), Influence of diurnal heating on stratification and residual circulation of Georges Bank, J. Geophys. Res., 108(C11), doi:10.1029/2001JC001245. 137 Chen, C., H. Liu, and R. C. Beardsley (2003b), An unstructured grid, finite-volume, threedimensional, primitive equations ocean model: application to coastal ocean and estuaries, J. Atmospheric Ocean. Technol., 20(1), 159–186. Chen, C., R. Beardsley, and G. Cowles (2006), An Unstructured Grid, Finite-Volume Coastal Ocean Model (FVCOM) System, Oceanography, 19(1), 78–89, doi:10.5670/oceanog.2006.92. Chikita, K., S. P. Joshi, J. Jha, and H. Hasegawa (2000), Hydrological and thermal regimes in a supra-glacial lake: Imja, Khumbu, Nepal Himalaya, Hydrol. Sci. J., 45(4), 507–521. Combes, S. (2003), Protecting freshwater ecosystems in the face of global climate change, Buy. Time User’s Man. Build. Resist. Resil. Clim. Change Nat. Syst. WWF Clim. Change Program. Comte, L., L. Buisson, M. Daufresne, and G. Grenouillet (2013), Climate-induced changes in the distribution of freshwater fish: observed and predicted trends: Climate change and freshwater fish, Freshw. Biol., 58(4), 625–639, doi:10.1111/fwb.12081. Constantz, J. (2008), Heat as a tracer to determine streambed water exchanges, Water Resour. Res., 44(4), doi:10.1029/2008WR006996. Covar, A. P. (1976), Selecting the Proper Reaeration Coefficient for Use in Water Quality Models, in Proceedings of the EPA Conference on Environmental Modeling and Simulation, Cincinnati, Ohio. Crumpton, W. G., T. M. Isenhart, and P. D. Mitchell (1992), Nitrate and organic N analyses with second-derivative spectroscopy, Limnol. Oceanogr., 37(4), 907. Dalin, C., Y. Wada, T. Kastner, and M. J. Puma (2017), Groundwater depletion embedded in international food trade, Nature, 543(7647), 700–704, doi:10.1038/nature21403. Dang, C., and H. Radha (2014), Single Image Super Resolution via Manifold Approximation, ArXiv Prepr. ArXiv14103349. Dang, C., and H. Radha (2015), Fast Image Super Resolution via Selective Manifold Learning of High Resolution Patches, Québec City, Canada. Dang, C., M. Aghagolzadeh, and H. Radha (2014), Image Super-Resolution via Local SelfLearning Manifold Approximation, IEEE Signal Process. Lett., 21(10), 1245–1249, doi:10.1109/LSP.2014.2332118. Dang, C. T., M. Aghagolzadeh, A. A. Moghadam, and H. Radha (2013), Single image super resolution via manifold linear approximation using sparse subspace clustering, in Global Conference on Signal and Information Processing (GlobalSIP), 2013 IEEE, pp. 949–952, IEEE, Austin, Texas, U.S.A. 138 Denman, K. L., and A. E. Gargett (1983), Time and space scales of vertical mixing and advection of phytoplankton in the upper ocean, Oceanography, 28(5). Dingman, S. L. (2002), Physical hydrology, 2nd ed., Prentice Hall, Upper Saddle River, N.J. Edson, J. B., V. Jampana, R. A. Weller, S. P. Bigorre, A. J. Plueddemann, C. W. Fairall, S. D. Miller, L. Mahrt, D. Vickers, and H. Hersbach (2013), On the Exchange of Momentum over the Open Ocean, J. Phys. Oceanogr., 43(8), 1589–1610, doi:10.1175/JPO-D-120173.1. Edwards, W. J., C. R. Rehmann, E. McDonald, and D. A. Culver (2005), The impact of a benthic filter feeder: limitations imposed by physical transport of algae to the benthos, Can. J. Fish. Aquat. Sci., 62(1), 205–214, doi:10.1139/f04-188. Elston, R., and G. S. L. Wood (1983), Pathogenesis of vibriosis in cultured juvenile red abalone, Haliotis rufescens Swainson, J. Fish Dis., 6(2), 111–128, doi:10.1111/j.13652761.1983.tb00059.x. Evensen, G. (2007), Data assimilation: the ensemble Kalman filter, Springer, Berlin; New York. Fairall, C. W., E. F. Bradley, D. P. Rogers, J. B. Edson, and G. S. Young (1996), Bulk parameterization of air-sea fluxes for Tropical Ocean-Global Atmosphere Coupled-Ocean Atmosphere Response Experiment, J. Geophys. Res. Oceans, 101(C2), 3747–3764, doi:10.1029/95JC03205. Farrand, W. R. (1982), Quaternary Geology of Michigan, Michigan Department of Environmental Quality, Geological Survey Division, Lansing, MI. Foreman, M. G. G., P. Czajko, D. J. Stucchi, and M. Guo (2009), A finite volume model simulation for the Broughton Archipelago, Canada, Ocean Model., 30(1), 29–47, doi:10.1016/j.ocemod.2009.05.009. Galperin, B., L. H. Kantha, S. Hassid, and A. Rosati (1988), A Quasi-equilibrium Turbulent Energy Model for Geophysical Flows, J. Atmospheric Sci., 45(1), 55–62, doi:10.1175/1520-0469(1988)045<0055:AQETEM>2.0.CO;2. Golub, G., and C. F. Van Loan (2013), Matrix Computations, 4th ed., Baltimore: Johns Hopkins. Gorman, G. J., M. D. Piggott, C. C. Pain, C. R. E. de Oliveira, A. P. Umpleby, and A. J. H. Goddard (2006), Optimisation based bathymetry approximation through constrained unstructured mesh adaptivity, Ocean Model., 12(3–4), 436–452, doi:10.1016/j.ocemod.2005.09.004. Goto, D., K. Lindelof, D. Fanslow, S. Ludsin, S. Pothoven, J. Roberts, H. Vanderploeg, A. Wilson, and T. Höök (2012), Indirect consequences of hypolimnetic hypoxia on zooplankton growth in a large eutrophic lake, Aquat. Biol., 16(3), 217–227, doi:10.3354/ab00442. Guillemin, V., and A. Pollack (2010), Differential topology, AMS Chelsea Pub, American Mathematical Society, Providence, R.I. 139 Gurrieri, J. T., and G. Furniss (2004), Estimation of groundwater exchange in alpine lakes using non-steady mass-balance methods, J. Hydrol., 297(1–4), 187–208, doi:10.1016/j.jhydrol.2004.04.021. Haidvogel, D. B., and A. Beckmann (1999), Numerical Ocean Circulation Modeling, Imp. Coll. Press. Haidvogel, D. B., H. G. Arango, K. Hedstrom, A. Beckmann, P. Malanotte-Rizzoli, and A. F. Shchepetkin (2000), Model evaluation experiments in the North Atlantic Basin: simulations in nonlinear terrain-following coordinates, Dyn. Atmospheres Oceans, 32(3), 239–281. Hampton, S. E., L. R. Izmest’Eva, M. V. Moore, S. L. Katz, B. Dennis, and E. A. Silow (2008), Sixty years of environmental change in the world’s largest freshwater lake - Lake Baikal, Siberia, Glob. Change Biol., 14(8), 1947–1958, doi:10.1111/j.1365-2486.2008.01616.x. Harris, J. O., C. M. Burke, S. J. Edwards, and D. R. Johns (2005), Effects of oxygen supersaturation and temperature on juvenile greenlip, Haliotis laevigata Donovan, and blacklip, Haliotis rubra Leach, abalone, Aquac. Res., 36(14), 1400–1407, doi:10.1111/j.13652109.2005.01360.x. Hecky, R. E., P. Campbell, and L. L. Hendzel (1993), The stoichiometry of carbon, nitrogen, and phosphorus in particulate matter of lakes and oceans, Limnol. Oceanogr., 38(4), 709–724, doi:10.4319/lo.1993.38.4.0709. Heiskanen, J. J. et al. (2015), Effects of water clarity on lake stratification and lake-atmosphere heat exchange, J. Geophys. Res. Atmospheres, 120(15), 2014JD022938, doi:10.1002/2014JD022938. Hondzo, M., and H. G. Stefan (1993), Lake Water Temperature Simulation Model, J. Hydraul. Eng., 119(11), 1251–1273, doi:10.1061/(ASCE)0733-9429(1993)119:11(1251). Hondzo, M., C. R. Ellis, and H. G. Stefan (1991), Vertical diffusion in small stratified lake: Data and error analysis, J. Hydraul. Eng., 117(10), 1352–1369. Horst, G. P., O. Sarnelle, J. D. White, S. K. Hamilton, R. B. Kaul, and J. D. Bressie (2014), Nitrogen availability increases the toxin quota of a harmful cyanobacterium, Microcystis aeruginosa, Water Res., 54, 188–198, doi:10.1016/j.watres.2014.01.063. Jackett, D. R., and T. J. McDougall (1995), Minimal Adjustment of Hydrographic Profiles to Achieve Static Stability, J. Atmospheric Ocean. Technol., 12(2), 381–389, doi:10.1175/1520-0426(1995)012<0381:MAOHPT>2.0.CO;2. Jiang, L., and X. Fang (2016), Simulation and Validation of Cisco Lethal Conditions in Minnesota Lakes under Past and Future Climate Scenarios Using Constant Survival Limits, Water, 8(7), 279, doi:10.3390/w8070279. Jiang, L., M. Xia, S. A. Ludsin, E. S. Rutherford, D. M. Mason, J. Marin Jarrin, and K. L. Pangle (2015), Biophysical modeling assessment of the drivers for plankton dynamics in 140 dreissenid-colonized western Lake doi:10.1016/j.ecolmodel.2015.04.004. Erie, Ecol. Model., 308, 18–33, Jiménez, P. A., and J. Dudhia (2012), Improving the Representation of Resolved and Unresolved Topographic Effects on Surface Wind in the WRF Model, J. Appl. Meteorol. Climatol., 51(2), 300–316, doi:10.1175/JAMC-D-11-084.1. Joodaki, G., J. Wahr, and S. Swenson (2014), Estimating the human contribution to groundwater depletion in the Middle East, from GRACE data, land surface models, and well observations, Water Resour. Res., 50(3), 2679–2692, doi:10.1002/2013WR014633. Kaushal, S. S., G. E. Likens, N. A. Jaworski, M. L. Pace, A. M. Sides, D. Seekell, K. T. Belt, D. H. Secor, and R. L. Wingate (2010), Rising stream and river temperatures in the United States, Front. Ecol. Environ., 8(9), 461–466, doi:10.1890/090037. Kettle, A. J., C. Hughes, G. A. Unazi, L. Birch, H. Mohie-El-Din, and M. R. Jones (2012), Role of groundwater exchange on the energy budget and seasonal stratification of a shallow temperate lake, J. Hydrol., 470–471, 12–27, doi:10.1016/j.jhydrol.2012.07.004. Kinsman-Costello, L. E., J. M. O’Brien, and S. K. Hamilton (2015), Natural stressors in uncontaminated sediments of shallow freshwaters: The prevalence of sulfide, ammonia, and reduced iron: Sulfide, ammonia, and iron in shallow freshwater sediments, Environ. Toxicol. Chem., 34(3), 467–479, doi:10.1002/etc.2801. Konikow, L. F. (2013), Groundwater Depletion in the United States (1900–2008), U.S. Geological Survey Scientific Investigations Report 2013−5079, U.S. Geological Survey, Reston, Virginia, United States. Kraus, E. B., and J. A. Businger (1995), Atmosphere-Ocean Interaction, Oxford University Press (US), Cary, US. Lane, P. A. (1979), Vertebrate and invertebrate predation intensity on freshwater zooplankton communities, Nature, 280(5721), 391–393, doi:10.1038/280391a0. Lane, P. V. Z., L. Llinás, S. L. Smith, and D. Pilz (2008), Zooplankton distribution in the western Arctic during summer 2002: Hydrographic habitats and implications for food chain dynamics, J. Mar. Syst., 70(1–2), 97–133, doi:10.1016/j.jmarsys.2007.04.001. Langston, G., M. Hayashi, and J. W. Roy (2013), Quantifying groundwater-surface water interactions in a proglacial moraine using heat and solute tracers: GW-SW Interaction in Proglacial Moraine, Water Resour. Res., 49(9), 5411–5426, doi:10.1002/wrcr.20372. Lazzaro, D., and L. B. Montefusco (2002), Radial basis functions for the multivariate interpolation of large scattered data sets, J. Comput. Appl. Math., 140(1–2), 521–536, doi:10.1016/S0377-0427(01)00485-X. Li, J., and A. D. Heap (2008), A review of spatial interpolation methods for environmental scientists, Record 2008/23, Geoscience Australia, Canberra. 141 Li, L., and P. Revesz (2004), Interpolation methods for spatio-temporal geographic data, Comput. Environ. Urban Syst., 28(3), 201–227, doi:10.1016/S0198-9715(03)00018-8. Li, L., T. Losser, C. Yorke, and R. Piltner (2014a), Fast Inverse Distance Weighting-Based Spatiotemporal Interpolation: A Web-Based Application of Interpolating Daily Fine Particulate Matter PM2:5 in the Contiguous U.S. Using Parallel Programming and k-d Tree, Int. J. Environ. Res. Public. Health, 11(9), 9101–9141, doi:10.3390/ijerph110909101. Li, M., L. Zhong, and W. C. Boicourt (2005), Simulations of Chesapeake Bay estuary: Sensitivity to turbulence mixing parameterizations and comparison with observations, J. Geophys. Res., 110(C12), doi:10.1029/2004JC002585. Li, R. et al. (2014b), Observed wintertime tidal and subtidal currents over the continental shelf in the northern South China Sea, J. Geophys. Res. Oceans, 119(8), 5289–5310, doi:10.1002/2014JC009931. Lucas, A. et al. (2016), Adrift Upon a Salinity-Stratified Sea: A View of Upper-Ocean Processes in the Bay of Bengal During the Southwest Monsoon, Oceanography, 29(2), 134–145, doi:10.5670/oceanog.2016.46. Ludwig, D., S. Carpenter, and W. Brock (2003), Optimal phosphorus loading for a potentially eutrophic lake, Ecol. Appl., 13(4), 1135–1152. Luo, L., J. Wang, D. J. Schwab, H. Vanderploeg, G. Leshkevich, X. Bai, H. Hu, and D. Wang (2012), Simulating the 1998 spring bloom in Lake Michigan using a coupled physicalbiological model, J. Geophys. Res. Oceans, 117(C10), n/a-n/a, doi:10.1029/2012JC008216. Luo, W., M. C. Taylor, and S. R. Parker (2008), A comparison of spatial interpolation methods to estimate continuous wind speed surfaces using irregularly distributed data from England and Wales, Int. J. Climatol., 28(7), 947–959, doi:10.1002/joc.1583. Ma, Y., P. Niyogi, G. Sapiro, and R. Vidal (2011), Dimensionality reduction via subspace and submanifold learning, IEEE Signal Process. Mag., 28(2), 14–126, doi:10.1109/MSP.2010.940005. van der Maaten, L. J., E. O. Postma, and H. J. van den Herik (2009), Dimensionality reduction: A comparative review, J. Mach. Learn. Res., 10(1–41), 66–71. MacEachren, A. M., and J. V. Davidson (1987), Sampling and isometric mapping of continuous geographic surfaces, Am. Cartogr., 14(4), 299–320. MacKay, M. D. et al. (2009), Modeling lakes and reservoirs in the climate system, Limnol. Oceanogr., 54(6), 2315. Madala, R. V., and S. A. Piacseki (1977), A semi-implicit numerical model for baroclinic oceans, J. Comput. Phys., 23(2), 167–178, doi:10.1016/0021-9991(77)90119-X. 142 Magnuson, J. J. (2000), Historical Trends in Lake and River Ice Cover in the Northern Hemisphere, Science, 289(5485), 1743–1746, doi:10.1126/science.289.5485.1743. Markfort, C. D., A. L. S. Perez, J. W. Thill, D. A. Jaster, F. Porté-Agel, and H. G. Stefan (2010), Wind sheltering of a lake by a tree canopy or bluff topography, Water Resour. Res., 46(3), W03530, doi:10.1029/2009WR007759. Martinho, A. S., and M. L. Batteen (2006), On reducing the slope parameter in terrain-following numerical ocean models, Ocean Model., 13(2), 166–175, doi:10.1016/j.ocemod.2006.01.003. Mellard, J. P., K. Yoshiyama, E. Litchman, and C. A. Klausmeier (2011), The vertical distribution of phytoplankton in stratified water columns, J. Theor. Biol., 269(1), 16–30, doi:10.1016/j.jtbi.2010.09.041. Mellor, G. L., and T. Yamada (1982), Development of a turbulence closure model for geophysical fluid problems, Rev. Geophys., 20(4), 851–875, doi:10.1029/RG020i004p00851. Mellor, G. L., T. Ezer, and L.-Y. Oey (1994), The Pressure Gradient Conundrum of Sigma Coordinate Ocean Models, J. Atmospheric Ocean. Technol., 11(4), 1126–1134, doi:10.1175/1520-0426(1994)011<1126:TPGCOS>2.0.CO;2. Mellor, G. L., L. Y. Oey, and T. Ezer (1998), Sigma coordinate pressure gradient errors and the seamount problem, J. Atmospheric Ocean. Technol., 15(5), 1122–1131. Menberg, K., P. Blum, B. L. Kurylyk, and P. Bayer (2014), Observed groundwater temperature response to recent climate change, Hydrol. Earth Syst. Sci., 18(11), 4453–4466, doi:10.5194/hess-18-4453-2014. Mendoza-Sanchez, I., M. S. Phanikumar, J. Niu, J. R. Masoner, I. M. Cozzarelli, and J. T. Mcguire (2013), Quantifying wetland-aquifer interactions in a humid subtropical climate region: An integrated approach, J. Hydrol., 498, 237–253. Merwade, V. (2009), Effect of spatial trends on interpolation of river bathymetry, J. Hydrol., 371(1–4), 169–181, doi:10.1016/j.jhydrol.2009.03.026. Monaghan, G. W., D. W. Forstat, and H. Q. Sorenson (1983), Selected geologic maps of Kalamazoo County, Michigan: County geologic map series, Michigan Geologic Survey Division, Map. Moradkhani, H., S. Sorooshian, H. V. Gupta, and P. R. Houser (2005), Dual state–parameter estimation of hydrological models using ensemble Kalman filter, Adv. Water Resour., 28(2), 135–147, doi:10.1016/j.advwatres.2004.09.002. Mordohai, P., and G. Medioni (2010), Dimensionality estimation, manifold learning and function approximation using tensor voting, J. Mach. Learn. Res., 11, 411–450. 143 Moss, B. (1972), Studies on Gull Lake, Michigan: I. Seasonal and depth distribution of phytoplankton, Freshw. Biol., 2(4), 289–307. Nguyen, T. D., P. Thupaki, E. J. Anderson, and M. S. Phanikumar (2014), Summer circulation and exchange in the Saginaw Bay-Lake Huron system, J. Geophys. Res. Oceans, 119(4), 2713– 2734, doi:10.1002/2014JC009828. Niu, J., C. Shen, S.-G. Li, and M. S. Phanikumar (2014), Quantifying storage changes in regional Great Lakes watersheds using a coupled subsurface-land surface process model and GRACE, MODIS products, Water Resour. Res., 50(9), 7359–7377, doi:10.1002/2014WR015589. Niu, Q., M. Xia, E. S. Rutherford, D. M. Mason, E. J. Anderson, and D. J. Schwab (2015), Investigation of interbasin exchange and interannual variability in Lake Erie using an unstructured-grid hydrodynamic model, J. Geophys. Res. Oceans, 120(3), 2212–2232, doi:10.1002/2014JC010457. Oleson, K. W. et al. (2010), Technical description of version 4.0 of the Community Land Model (CLM), NCAR Technical Note, National Center for Atmospheric Research, Boulder, Colorado. O’Reilly, C. M., S. R. Alin, P.-D. Plisnier, A. S. Cohen, and B. A. McKee (2003), Climate change decreases aquatic ecosystem productivity of Lake Tanganyika, Africa, Nature, 424(6950), 766–768, doi:10.1038/nature01833. Parkinson, C. L., and W. M. Washington (1979), A large-scale numerical model of sea ice, J. Geophys. Res. Oceans, 84(C1), 311–337, doi:10.1029/JC084iC01p00311. Pathiraja, S., L. Marshall, A. Sharma, and H. Moradkhani (2016), Hydrologic modeling in dynamic catchments: A data assimilation approach, Water Resour. Res., n/a-n/a, doi:10.1002/2015WR017192. Paulson, C. A., and J. J. Simpson (1977), Irradiance Measurements in the Upper Ocean, J. Phys. Oceanogr., 7(6), 952–956, doi:10.1175/1520-0485(1977)007<0952:IMITUO>2.0.CO;2. Perry, L., and C. J. D. Brown (1942), A fisheries survey of Gull Lake, Kalamazoo and Barry counties, Fisheries Research Report No. 725, Michigan Department of Natural Resources, Fisheries Division, Ann Arbor, Michigan, USA. Pokhrel, Y. N., S. Koirala, P. J.-F. Yeh, N. Hanasaki, L. Longuevergne, S. Kanae, and T. Oki (2015), Incorporation of groundwater pumping in a global Land Surface Model with the representation of human impacts, Water Resour. Res., 51(1), 78–96, doi:10.1002/2014WR015602. Press, W. H. (Ed.) (2007), Numerical recipes: the art of scientific computing, 3rd ed., Cambridge University Press, Cambridge, UK ; New York. 144 Rasmussen, A. H., M. Hondzo, and H. G. Stefan (1995), A TEST OF SEVERAL EVAPORATION EQUATIONS FOR WATER TEMPERATURE SIMULATIONS IN LAKES, J. Am. Water Resour. Assoc., 31(6), 1023–1028, doi:10.1111/j.1752-1688.1995.tb03418.x. Renka, R. J., and R. Brown (1999), Algorithm 792: Accuracy Test of ACM Algorithms for Interpolation of Scattered Data in the Plane, ACM Trans Math Softw, 25(1), 78–94, doi:10.1145/305658.305745. Rodell, M., I. Velicogna, and J. S. Famiglietti (2009), Satellite-based estimates of groundwater depletion in India, Nature, 460(7258), 999–1002, doi:10.1038/nature08238. Rodi, W. (1987), Examples of calculation methods for flow and mixing in stratified fluids, J. Geophys. Res. Oceans, 92(C5), 5305–5328. Rowe, M. D., E. J. Anderson, J. Wang, and H. A. Vanderploeg (2015), Modeling the effect of invasive quagga mussels on the spring phytoplankton bloom in Lake Michigan, J. Gt. Lakes Res., doi:10.1016/j.jglr.2014.12.018. Roweis, S. T., and L. K. Saul (2000), Nonlinear dimensionality reduction by locally linear embedding, Science, 290(5500), 2323–2326, doi:10.1126/science.290.5500.2323. Rueda, F. J., S. G. Schladow, S. G. Monismith, and M. T. Stacey (2005), On the effects of topography on wind and the generation of currents in a large multi-basin lake, Hydrobiologia, 532(1–3), 139–151. Rühland, K. M., A. M. Paterson, and J. P. Smol (2015), Lake diatom responses to warming: reviewing the evidence, J. Paleolimnol., 54(1), 1–35, doi:10.1007/s10933-015-9837-3. Safaie, A., A. Wendzel, Z. Ge, M. B. Nevers, R. L. Whitman, S. R. Corsi, and M. S. Phanikumar (2016), Comparative Evaluation of Statistical and Mechanistic Models of Escherichia coli at Beaches in Southern Lake Michigan, Environ. Sci. Technol., 50(5), 2442–2449, doi:10.1021/acs.est.5b05378. Safaie, A., C. Dang, H. Qiu, H. Radha, and M. S. Phanikumar (2017a), Manifold methods for assimilating geophysical and meteorological data in Earth system models and their components, J. Hydrol., 544, 383–396, doi:10.1016/j.jhydrol.2016.11.009. Safaie, A., E. Litchman, and M. S. Phanikumar (2017b), Evaluating the Role of Groundwater in Circulation and Thermal Structure within a Deep Inland Lake, Advances in Water Resources, doi:10.1016/j.advwatres.2017.08.002. Sarnelle, O., A. E. Wilson, S. K. Hamilton, L. B. Knoll, and D. F. Raikow (2005), Complex interactions between the zebra mussel, Dreissena polymorpha, and the harmful phytoplankter, Microcystis aeruginosa, Limnol. Oceanogr., 50(3), 896–904, doi:10.4319/lo.2005.50.3.0896. Schindler, D. W. (2009), Lakes as sentinels and integrators for the effects of climate change on watersheds, airsheds, and landscapes, Limnol. Oceanogr., 54(6), 2349. 145 Schladow, S. G., and D. P. Hamilton (1997), Prediction of water quality in lakes and reservoirs: Part II - Model calibration, sensitivity analysis and application, Ecol. Model., 96(1–3), 111–123, doi:10.1016/S0304-3800(96)00063-4. Schneider, P., and S. J. Hook (2010), Space observations of inland water bodies show rapid surface warming since 1985, Geophys. Res. Lett., 37(22), n/a-n/a, doi:10.1029/2010GL045059. Schumann, U., and T. Gerz (1995), Turbulent Mixing in Stably Stratified Shear Flows, J. Appl. Meteorol., 34(1), 33–48, doi:10.1175/1520-0450-34.1.33. Schuster, P. F., M. M. Reddy, J. W. LaBaugh, R. S. Parkhurst, D. O. Rosenberry, T. C. Winter, R. C. Antweiler, and W. E. Dean (2003), Characterization of lake water and ground water movement in the littoral zone of Williams Lake, a closed-basin lake in north central Minnesota, Hydrol. Process., 17(4), 823–838, doi:10.1002/hyp.1211. Schwab, D. J., and D. Beletsky (1998), Lake Michigan Mass Balance Study: Hydrodynamic modeling project, Great Lakes Environmental Research Laboratory, Ann Arbor, MI. Schwab, D. J., and J. A. Morton (1984), Estimation of Overlake Wind Speed from Overland Wind Speed: A Comparison of Three Methods, J. Gt. Lakes Res., 10(1), 68–72, doi:10.1016/S0380-1330(84)71808-9. Shigesada, N., and A. Okubo (1981), Analysis of the self-shading effect on algal vertical distribution in natural waters, J. Math. Biol., 12(3), 311–326. Shore, J. A. (2009), Modelling the circulation and exchange of Kingston Basin and Lake Ontario with FVCOM, Ocean Model., 30(2–3), 106–114, doi:10.1016/j.ocemod.2009.06.007. Sikirić, M. D., I. Janeković, and M. Kuzmić (2009), A new approach to bathymetry smoothing in sigma-coordinate ocean models, Ocean Model., 29(2), 128–136, doi:10.1016/j.ocemod.2009.03.009. Šiljeg, A., S. Lozić, and S. Šiljeg (2015), A comparison of interpolation methods on the basis of data obtained from a bathymetric survey of Lake Vrana, Croatia, Hydrol. Earth Syst. Sci., 19(8), 3653–3666, doi:10.5194/hess-19-3653-2015. Simons, T. J. (1974), Verification of Numerical Models of Lake Ontario: Part I. Circulation in Spring and Early Summer, J. Phys. Oceanogr., 4(4), 507–523, doi:10.1175/15200485(1974)004<0507:VONMOL>2.0.CO;2. Simpson, J. J., and T. D. Dickey (1981a), Alternative Parameterizations of Downward Irradiance and Their Dynamical Significance, J. Phys. Oceanogr., 11(6), 876–882, doi:10.1175/15200485(1981)011<0876:APODIA>2.0.CO;2. Simpson, J. J., and T. D. Dickey (1981b), The Relationship between Downward Irradiance and Upper Ocean Structure, J. Phys. Oceanogr., 11(3), 309–323, doi:10.1175/15200485(1981)011<0309:TRBDIA>2.0.CO;2. 146 Sommaruga, R., R. Psenner, E. Schafferer, K. A. Koinig, and S. Sommaruga-Wograth (1999), Dissolved Organic Carbon Concentration and Phytoplankton Biomass in High-Mountain Lakes of the Austrian Alps: Potential Effect of Climatic Warming on UV Underwater Attenuation, Arct. Antarct. Alp. Res., 31(3), 247, doi:10.2307/1552253. Sommaruga-Wögrath, S., K. A. Koinig, R. Schmidt, R. Sommaruga, R. Tessadri, and R. Psenner (1997), Temperature effects on the acidity of remote alpine lakes, Nature, 387(6628), 64– 67, doi:10.1038/387064a0. Staffell, I., and S. Pfenninger (2016), Using bias-corrected reanalysis to simulate current and future wind power output, Energy, 114, 1224–1239, doi:10.1016/j.energy.2016.08.068. Stepanenko, V., K. D. Jöhnk, E. Machulskaya, M. Perroud, Z. Subin, A. Nordbo, I. Mammarella, and D. Mironov (2014), Simulation of surface energy fluxes and stratification of a small boreal lake by a set of one-dimensional models, Tellus Dyn. Meteorol. Oceanogr., 66(1), 21389, doi:10.3402/tellusa.v66.21389. Sterman, J. (2000), Business dynamics: systems thinking and modeling for a complex world, Irwin/McGraw-Hill, Boston. Suparta, W., and R. Rahman (2016), Spatial interpolation of GPS PWV and meteorological variables over the west coast of Peninsular Malaysia during 2013 Klang Valley Flash Flood, Atmospheric Res., 168, 205–219, doi:10.1016/j.atmosres.2015.09.023. Tague, D. F. (1977), The hydrologic and total phosphorus budgets of Gull Lake, Michigan, Michigan State University, East Lansing, Michigan, United States. Tenenbaum, J. B., V. De Silva, and J. C. Langford (2000), A global geometric framework for nonlinear dimensionality reduction, Science, 290(5500), 2319–2323. 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(3), 1606–1617, doi:10.1002/jgrc.20136. Turner, J. V., and L. R. Townley (2006), Determination of groundwater flow-through regimes of shallow lakes and wetlands from numerical analysis of stable isotope and chloride tracer distribution patterns, J. Hydrol., 320(3–4), 451–483, doi:10.1016/j.jhydrol.2005.07.050. Umaña, G. (2014), Ten years of limnological monitoring of a modified natural lake in the tropics: Cote Lake, Costa Rica, Rev. Biol. Trop., 62(2), 576–578. Umlauf, L., and H. Burchard (2003), A generic length-scale equation for geophysical turbulence models, J. Mar. Res., 61(2), 235–265, doi:10.1357/002224003322005087. von Rohden, C., J. Ilmberger, and B. Boehrer (2009), Assessing groundwater coupling and vertical exchange in a meromictic mining lake with an SF6-tracer experiment, J. Hydrol., 372(1– 4), 102–108, doi:10.1016/j.jhydrol.2009.04.004. 147 Wada, Y., L. P. H. van Beek, C. M. van Kempen, J. W. T. M. Reckman, S. Vasak, and M. F. P. Bierkens (2010), Global depletion of groundwater resources: GLOBAL GROUNDWATER DEPLETION, Geophys. Res. Lett., 37(20), n/a-n/a, doi:10.1029/2010GL044571. Webster, K. E., T. K. Kratz, C. J. Bowser, J. J. Magnuson, and W. J. Rose (1996), The influence of landscape position on lake chemical responses to drought in northern Wisconsin, Limnol. Oceanogr., 41(5), 977–984. Weitkamp, D. E., and M. Katz (1980), A Review of Dissolved Gas Supersaturation Literature, Trans. Am. Fish. Soc., 109(6), 659–702, doi:10.1577/15488659(1980)109<659:ARODGS>2.0.CO;2. White, J., and S. K. Hamilton (2014), Gull Lake watershed monitoring Program 2014 annual report, The Gull Lake Quality Organization, Hickory Corners, MI. White, J. D., S. K. Hamilton, O. Sarnelle, and K. Tierney (2015), Heat-induced mass mortality of invasive zebra mussels ( Dreissena polymorpha ) at sublethal water temperatures, Can. J. Fish. Aquat. Sci., 72(8), 1221–1229, doi:10.1139/cjfas-2015-0064. Williamson, C. E., R. S. Stemberger, D. P. Morris, T. M. Frost, and S. G. Paulsen (1996), Ultraviolet radiation in North American lakes: attenuation estimates from DOC measurements and implications for plankton communities, Limnol. Oceanogr., 41(5), 1024–1034. Wilson, A. E., and O. Sarnelle (2002), Relationship between zebra mussel biomass and total phosphorus in European and North American lakes, Arch. Hydrobiol., 153(2). Wilson, M. C., J. A. Shore, and Y. R. Rao (2013), Sensitivity of the Simulated Kingston Basin— Lake Ontario Summer Temperature Profile using FVCOM, Atmosphere-Ocean, 51(3), 319–331, doi:10.1080/07055900.2013.800017. Wilson, S. E. (1987), Bedrock Geology Map of Michigan, Michigan Department of Environmental Quality, Geological Survey Division, Lansing, MI. Winslow, L. A., J. S. Read, G. J. A. Hansen, and P. C. Hanson (2015), Small lakes show muted climate change signal in deepwater temperatures, Geophys. Res. Lett., 42(2), 2014GL062325, doi:10.1002/2014GL062325. Wollschläger, U., J. Ilmberger, M. Isenbeck-Schröter, A. M. Kreuzer, C. von Rohden, K. Roth, and W. Schäfer (2007), Coupling of groundwater and surface water at Lake Willersinnweiher: Groundwater modeling and tracer studies, Aquat. Sci., 69(1), 138–152, doi:10.1007/s00027-006-0825-6. Woolway, R. I. et al. (2016), Diel surface temperature range scales with lake size, PloS One, 11(3), e0152466. 148 Xue, P., D. J. Schwab, and S. Hu (2015), An investigation of the thermal response to meteorological forcing in a hydrodynamic model of Lake Superior, J. Geophys. Res. Oceans, 120(7), 5233–5253, doi:10.1002/2015JC010740. Yacobi, Y. Z., and T. Zohary (2010), Carbon:chlorophyll a ratio, assimilation numbers and turnover times of Lake Kinneret phytoplankton, Hydrobiologia, 639(1), 185–196, doi:10.1007/s10750-009-0023-3. Yan, Y., F. Xiao, and Y. Du (2014), Construction of lake bathymetry from MODIS satellite data and GIS from 2003 to 2011, Chin. J. Oceanol. Limnol., 32(3), 720–731, doi:10.1007/s00343-014-3185-4. Yassuda, E. A., S. R. Davie, D. L. Mendelsohn, T. Isaji, and S. J. Peene (2000), Development of a waste load allocation model for the Charleston Harbor estuary, phase II: water quality, Estuar. Coast. Shelf Sci., 50(1), 99–107. Yoshiyama, K., J. P. Mellard, E. Litchman, and C. A. Klausmeier (2009), Phytoplankton Competition for Nutrients and Light in a Stratified Water Column, Am. Nat., 174(2), 190– 203, doi:10.1086/600113. Zhang, H., I. Mendoza-Sanchez, E. L. Miller, and L. M. Abriola (2016a), Manifold Regression Framework for Characterizing Source Zone Architecture, IEEE Trans. Geosci. Remote Sens., 54(1), 3–17, doi:10.1109/TGRS.2015.2448086. Zhang, Y., Z. Wu, M. Liu, J. He, K. Shi, Y. Zhou, M. Wang, and X. Liu (2015), Dissolved oxygen stratification and response to thermal structure and long-term climate change in a large and deep subtropical reservoir (Lake Qiandaohu, China), Water Res., 75, 249–258, doi:10.1016/j.watres.2015.02.052. Zhang, Y., Y. Su, Z. Liu, E. Jeppesen, J. Yu, and M. Jin (2016b), Geochemical records of anoxic water mass expansion in an oligotrophic alpine lake (Yunnan Province, SW China) in response to climate warming since the 1980s, The Holocene, 0959683616645948. Zheng, L., C. Chen, and F. Y. Zhang (2004), Development of water quality model in the Satilla River Estuary, Georgia, Ecol. Model., 178(3–4), 457–482, doi:10.1016/j.ecolmodel.2004.01.016. Zundel, A. K., and J. L. Jones (1996), An Integrated Surface Water Modeling System, International Association for Hydraulic Research HYDROINFORMATICS, Zurich, Switzerland. 149