Skip to main content

Full text of "Modeling the circulation and water quality in Charlotte Harbor Estuarine System, Florida"

See other formats


MODELING  THE  CIRCULATION  AND  WATER  QUALITY 
IN  CHARLOTTE  HARBOR  ESTUARINE  SYSTEM,  FLORIDA 


By 

KIJFN  PARK 


A  DISSERTATION  PRESENTED  TO  THE  GRADUATE  SCHOOL 

OF  THE  UNIVERSITY  OF  FLORIDA  IN  PARTIAL  FULFILLMENT 

OF  THE  REQUIREMENTS  FOR  THE  DEGREE  OF 

DOCTOR  OF  PHILOSOPHY 


UNIVERSITY  OF  FLORIDA 
2004 


ACKNOWLEDGMENTS 

I  would  like  to  thank  my  advisor,  Dr.  Y.  Peter  Sheng,  for  his  guidance,  support  and 
financial  assistance  throughout  my  study.  In  addition,  much  appreciation  is  owned  to  my 
other  committee  members,  Dr.  Robert  G.  Dean,  Dr.  Robert  J.  Thieke,  Dr.  Louis  H.  Motz  and 
Dr.  K.  Ramesh  Reddy,  for  their  review  of  my  dissertation. 

I  would  like  to  thank  the  South  Florida  Water  Management  District  and  Southwest 
florida  Water  Management  District  for  sponsoring  research  project  and  providing  data  for 
Charlotte  Harbor  estuarine  system.  I  would  like  to  express  my  thanks  to  Justin,  Jeffery, 
Yangfeng,  Taeyoon,  Jun,  Vadim,  and  Vladimir  whose  help  with  class,  research  and  writing 
this  dissertation.  Many  thanks  go  to  Becky,  Lucy,  Sonna,  Sidney,  Kim  and  Halen  for  making 
life  easier. 

I  would  like  to  dedicate  this  dissertation  to  my  parents  whose  love  and  support  made 
this  degree  possible.  Last,  but  not  least,  I  would  like  to  thank  my  wife,  Kyung-Mi,  who  have 
been  praying  for  me  to  be  faithful,  kind,  and  honest  all  the  time. 


11 


TABLE  OF  CONTENTS 

Page 

ACKNOWLEDGMENTS » 

LIST  OF  TABLES vi 

LIST  OF  FIGURES  ix 

ABSTRACT xvil 

CHAPTER 

1  INTRODUCTION 1 

2  CHARLOTTE  HARBOR  CHARACTERIZATION   9 

2.1  Climate U 

2.2  Hydrodynamics   1 1 

2.2.1  Tidal  Stage,  Discharge  at  Inlet,  and  Tidal  Circulation  13 

2.2.2  Freshwater  Flow 14 

2.2.3  Salinity   14 

2.3  Water  Quality 15 

2.3.1  Nutrients 16 

2.3.2  Dissolved  Oxygen 18 

2.3.3  Phytoplankton 19 

2.4  Sediment    20 

2.5  Light  Environment 21 

3  HYDRODYNAMIC  AND  SEDIMENT  TRANSPORT  MODEL 23 

3.1  Governing  Equation 24 

3.1.1  Hydrodynamic  Model 24 

3.1.2  Sediment  Transport  Model 26 

3.2  Boundary  and  Initial  Conditions    27 

3.2.1  Boundary  Conditions 27 

3.2.2  Initial  Conditions    30 

3.3  Heat-Flux  at  Air-Sea  Interface 30 

3.3.1  Short-Wave  Solar  Radiation 31 

3.3.2  Long-Wave  Solar  Radiation 32 

in 


3.3.3  Sensible  and  Latent  Heat  Fluxes 33 

4  WATER  QUALITY  MODEL 36 

4.1  Mathematical  Formulae 38 

4.2  Phytoplankton  Dynamics 40 

4.2.1  Modeling  Approach    40 

4.2.2  Relationship  between  Phytoplankton  and  Nutrients 42 

4.3  Nutrient  Dynamics 43 

4.4  Oxygen  Balance 45 

4.5  Effects  of  Temperature  and  Light  Intensity  on  Water  Quality  Model 59 

4.5.1  Temperature     59 

4.5.2  Light  intensity 61 

4.6  Light  Attenuation  Model 61 

4.7  Model  Parameters  and  Calibration  Procedures  69 

5  APPLICATION  OF  CIRCULATION  AND  TRANSPORT  MODEL 81 

5.1  A  High-Resolution  Curvilinear  Grid  for  Charlotte  Harbor  Estuarine  System  82 

5.2  Forcing  Mechanism  and  Boundary  Condition 86 

5.3  Simulations  for  1986  Hydrodynamics 94 

5.3.1  Sensitivity  and  Calibration  Simulations 99 

5.3.2  Results  of  the  1986  simulation  101 

5.4  Simulations  for  2000  Hydrodynamic 118 

5.4.1  Sensitivity  and  Calibration  Simulations 118 

5.4.2  Results  of  the  2000  Simulation 130 

5.4.3  Application  of  the  2000  hydrodynamic  Simulations    146 

6  APPLICATION  OF  WATER  QUALITY  MODEL 167 

6.1  Forcing  Mechanism  and  Boundary  Condition  for  Circulation    167 

6.2  Initial  and  Boundary  Condition  for  the  Water  Quality  Model 172 

6.3  Simulations  of  Water  Quality  in  1996 180 

6.3.1  Calibration 180 

6.3.2  Results  for  1996  Water  Quality  Simulations    191 

6.4  Simulations  of  Water  Quality  in  2000 212 

6.4.1  Verification 212 

6.4.2  Results  of  2000  Water  Quality  Simulations 215 

6.4.3  Application  of  2000  Water  Quality  Simulations 235 

7  CONCLUSION  AND  DISCUSSION 258 

APPENDIX 

A  FLOW  CHARTS  FOR  CH3DJMS    263 


IV 


B  DMENSIONLESS  EQUATIONS  IN  CURVILINEAR  BOUNDARY-FITTED  AND 
SIGMA  GRID   278 

C  COMPARISON  OF  WATER  QUALITY  MODELS 284 

D  NUMERICAL  SOLUTION  TECHNIQUES  FOR  WATER  QUALITY  MODEL  .  .  289 

E  NUTRIENT  DYNAMICS 294 

F  SEDIMENT  FLUX  MODEL 306 

G  MODEL  PERFORMANCE  TEST  WITH  PARALLEL  CH3D_IMS    309 

H  TIDAL  BENCH  MARKS  FOR  CHARLOTTE  HARBOR   313 

REFERENCES    314 

BIOGRAPHICAL  SKETCH 325 


LIST  OF  TABLES 
Table  E^e 

1.1  Component  models  of  the  CH3D-IMS 6 

2.1  Ratios  of  nitrogen  and  phosphorous  constituents 18 

3.1  Mean  latitudinal  values  of  the  coefficient  X 33 

4. 1  Average  values  of  oxygen  uptake  rates  of  bottom 54 

4.2  The  spectrum  of  incident  sunlight  data 64 

4.3  Coefficient  ranges  to  use  in  stand  along  light  model 67 

4.4  The  best  fit  light  model  coefficients  for  Charlotte  Harbor  estuarine  system 67 

4.5  Dixon  and  Gray's  model  coefficients  for  the  Charlotte  Harbor  estuarine  system..    .  67 

4.6  Temperature  adjustment  functions  for  water  quality  parameters 70 

4.7  Water  quality  parameters  related  to  conversion  rate 71 

4.8  Water  quality  parameters  related  to  phytoplankton  and  zooplankton 71 

4.9  Water  quality  parameters  in  the  nitrogen  dynamics 72 

4.10  Water  quality  parameters  in  the  phosphorous  dynamics 73 

4.11  Water  quality  parameters  in  the  oxygen  balance 74 

4.12  The  relationship  between  water  quality  parameters  and  model  constituents 79 

5.1  Descriptions  of  1986  and  2000  river  boundary  conditions  for  Charlotte  Harbor 

estuarine  system 87 

5.2  Locations  of  tidal  stage  and  velocity-salinity  measured  stations  by  USGS 96 


VI 


5.3  The  effect  of  removing  selected  boundary  conditions  on  the  accuracy  of  simulated 

water  level,  velocity  and  salinity  in  July  1986.  Value  shown  are  average  RMS 
differences  vs.  baseline  simulation  at  all  data  stations 99 

5.4  The  effect  of  varying  bottom  roughness,  z0,  on  the  accuracy  of  simulated  water  level, 

velocity  and  salinity  in  July  1986.  Value  shown  are  average  RMS  errors  at  all 
data  stations 100 

5.5  The  effect  of  varying  horizontal  diffusion,  AH,  on  the  accuracy  of  simulated  water 

level,  velocity  and  salinity  in  July  1986.  Value  shown  average  RMS  errors  at  all 
data  stations 101 

5.6  A  summary  of  boundary  conditions  and  model  parameters  used  in  the  1986 

simulation 102 

5.7  Calculated  RMS  errors  between  simulated  and  measured  water  level  for  1986 

simulation 102 

5.8  Calculated  RMS  errors  between  simulated  and  measured  current  velocity  for  1986 

simulation 104 

5.9  Calculated  RMS  errors  between  simulated  and  measured  salinity  for  1986 

simulation 105 

5.10  The  effect  of  horizontal  grid  resolution,  on  the  accuracy  of  simulated  water  level 

and  salinity.  Values  shown  are  average  RMS  errors  for  2000  calibration  at  all 
available  stations.  Values  shown  in  parenthesis  are  %  RMS  error  normalized  by 
maximum  values 121 

5.1 1  The  effect  of  vertical  grid  resolution,  on  the  accuracy  of  simulated  water  level  and 

salinity.  Values  shown  are  average  RMS  errors  for  2000  calibration  at  all 
available  stations.  Values  shown  in  parenthesis  are  %  RMS  error  normalized  by 
maximum  values 122 

5.12  The  effect  of  varying  bottom  roughness,  z0,  on  the  accuracy  of  simulated  water  level 

and  salinity  in  2000.  Values  shown  are  average  RMS  errors  all  data  stations.  .  123 

5.13  The  effect  of  varying  salinity  advection  scheme  on  the  accuracy  of  simulated  water 

level  and  salinity  in  2000.  Values  shown  are  average  RMS  errors  at  all  data 
stations 128 

5.14  The  effect  of  modifying  bathymetry  on  the  accuracy  of  simulated  water  level  and 

salinity  in  2000.  Values  shown  are  average  RMS  errors  for  all  data  stations.  .  129 


vn 


5. 15  A  summary  of  boundary  conditions  and  model  parameters  used  in  the  2000 

simulation ljyj 

5.16  Calculated  RMS  errors  between  simulated  and  measured  water  level  for  2000 

simulation "1 

5.17  Calculated  RMS  errors  between  simulated  and  measured  salinity  for  2000 

simulation "3 

5.18  The  effect  of  hydrologic  alternations  on  2000  water  level.  Value  shown  are 

average  RMS  differences  with  baseline  simulation  for  all  selected  stations.    . .  154 

5.19  The  effect  of  hydrologic  alternations  on  2000  salinity.  Value  shown  are  average 

RMS  differences  with  baseline  simulation  for  all  selected  stations 155 

6.1  Locations  of  water  quality  measured  stations 173 

6.2  Sediment  types  for  Charlotte  Harbor  water  quality  simulations 174 

6.3  Water  quality  parameters,  baseline  values,  and  ranges  used  in  the  sensitivity 

analysis 1" 

6.4  Sensitivity  analysis  results  in  RMS  difference  w.r.t.  baseline  for  1996  water  quality 

calibration  simulation 184 

6.5  The  water  quality  model  coefficients  used  for  the  Charlotte  Harbor  simulation.  .  .  185 

6.6  The  temporally  averaged  water  quality  species  concentrations  for  baseline  2000 

simulation 240 

6.7  Normalized  RMS  differences  of  water  quality  species  concentrations  at  1 1  stations 

during  April  to  June  2000,  showing  the  effect  of  no  causeway  islands 240 

6.8  Normalized  RMS  differences  of  water  quality  species  concentrations  at  11  stations 

during  April  to  June  2000,  showing  the  effect  of  no  ICW 241 

6.9  Normalized  RMS  differences  of  watei  quality  species  concentrations  at  1 1  stations 

during  April  to  June  2000,  showing  the  effect  of  no  causeway  islands  and  ICW241 

C.l  Comparison  of  water  quality  models 287 

G.  1  CPU  time  for  the  parallel,  shared  memory,  CH3D  procedures  on  SGI  origin 

platform 311 

H.l  Tidal  datum  referred  to  Mean  Low  Low  Water  (MLLW),  in  meter 313 

viii 


LIST  OF  FIGURES 

Figure                                                                                                                          Page 
1.1  Charlotte  Harbor  estuarine  system  and  its  subarea  boundaries 2 

2.1  Drainage  basins  of  Charlotte  Harbor  estuarine  system 10 

2.2  Seasonal  wind  pattern  in  Florida 12 

2.3  Average  monthly  concentration  of  dissolved  oxygen  in  upper  Charlotte  Harbor,  site 

CH-006,  1976-84 19 

4.1  The  connection  between  nitrogen,  phosphorous  and  carbon  cycle 44 

4.2  CBOD  cycle  and  DO  cycle 46 

4.3  Comparison  of  wind-dependent  re-aeration  formula   49 

4.4  The  relationship  between  POM  flux  and  SOD  flux  related  in  the  oxidation  and 

reduction  of  organic  matter  in  sediment  column     53 

4.5  Effect  of  dissolved  oxygen  on  sediment  consumption  and  SOD  release  57 

4.6  The  scatter  plots  for  kd(PAR)  during  calibration  period  with  best  fit  coefficients  .  .  68 

4.7  Systematic  calibration  procedure   80 

5.1  Boundary-fitted  grid  (92  x  129)  used  for  numerical  simulation  for  Charlotte  Harbor 

estuarine  system 84 

5.2  Bathymetry  in  the  boundary-fitted  grid  for  Charlotte  Harbor  estuarine  system  (92  x 

129)    85 

5.3  Tidal  forcing  and  river  discharges  for  1986  simulations  of  Charlotte  Harbor 

circulation 89 

5.4  Wind  velocity  for  1986  simulations  for  Charlotte  Harbor  circulation  90 

ix 


5.5  Tidal  forcing  and  river  discharges  for  2000  simulations  of  Charlotte  Harbor 

circulation 91 

5.6  Wind  velocity  for  2000  simulation  of  Charlotte  Harbor  circulation 92 

5.7  Air  temperature  for  2000  simulation  of  Charlotte  Harbor  circulation   93 

5.8  Locations  of  the  available  1986  water  level  and  discharge  measurement  stations  of 

USGS 97 

5.9  Locations  of  available  1986  velocity  and  salinity  measurement  stations  of  USGS   .  98 

5.10  Comparison  between  simulated  and  measured  water  level  in  July  1986  108 

5. 1 1  Comparison  between  simulated  and  measure  spectra  of  water  level  in  July  1986  1 10 

5.12  Comparison  between  simulated  and  measured  current  velocity  in  July  1986    ...  1 1 1 

5.13  Comparison  between  simulated  and  measured  salinity  in  July  1986  115 

5.14  Typical  flow  pattern  of  Charlotte  Harbor  estuarine  system  during  one  tidal  cycle  for 

August  6,  1986 H6 

5.15  The  29-day  residual  flow  and  salinity  for  Charlotte  Harbor  estuarine  system  during 

July  2  to  July  30,  1986 117 

5.16  Locations  of  the  available  2000  water  level  and  salinity  measured  stations  at 

Caloosahatchee  River  operated  by  SFWMD 119 

5.17  The  comparison  of  the  coarse  grid  (71x92)  and  the  fine  grid  (92x129)  for  Charlotte 

Harbor  estuarine  system 120 

5.18  A  comparison  between  simulated  and  measured  salinity  at  Shell  Point  using 

ultimate  QUICKEST,  QUICKEST,  and  upwind  advection  schemes 124 

5.19  A  comparison  between  simulated  and  measured  salinity  at  Fort  Myers  using 

ultimate  QUICKEST,  QUICKEST,  and  upwind  advection  schemes 125 

5.20  A  comparison  between  simulated  and  measured  salinity  at  BR31  using  ultimate 

QUICKEST,  QUICKEST,  and  upwind  advection  schemes 126 

5.21  Simulated  longitudinal-vertical  salinity  along  the  Caloosahatchee  River  at  slack- 

water  before  flood  tide  on  September  7,  2000 127 

5.22  Comparison  between  simulated  and  measured  water  level  in  2000 136 

x 


5.23  Comparison  between  simulated  and  measured  salinity  at  S79  in  2000 137 

5.24  Comparison  between  simulated  and  measured  salinity  at  BR31  in  2000 138 

5.25  Comparison  between  simulated  and  measured  salinity  at  Fort  Myers  in  2000.  .  .  139 

5.26  Comparison  between  simulated  and  measured  salinity  at  Shell  Point  in  2000.  .  .  140 

5.27  Comparison  between  simulated  and  measured  salinity  near  Sanibel  Causeway  in 

2000 141 

5  28  Comparison  between  simulated  and  measured  temperature  at  Fort  Myers  in  2000. 

142 

5.29  Typical  flow  pattern  of  San  Carlos  Bay  during  ebb  and  flood  tide  for  August,  7  on 

2000 143 

5.30  One-year  residual  flow  in  San  Carlos  Bay  in  2000 144 

5.31  One-year  residual  salinity  distribution  in  San  Carlos  Bay  in  2000 145 

5.32  The  locations  of  Sanibel  Causeway  and  IntraCoastal  Waterway  and  stations  for 

comparing  the  effect  of  hydrologic  alterations    148 

5.33  The  comparison  of  bathymetry  and  shoreline  for  each  hydrologic  alteration  case 

scenarios  which  are  Baseline,  the  absence  of  IntraCoastal  Waterway,  and  the 
absence  of  causeway l4y 

5.34  The  comparisons  of  water  level  for  three  cases  at  three  selected  stations:  ST05  (Pine 

Island  Sound),  ST08  (San  Carlos  Bay),  and  ST10  (Caloosahatchee  River  mouth). 
150 

5.35  The  comparisons  of  surface  and  bottom  salinity  for  three  cases  at  three  selected 

stations:  ST05  (Pine  Island  Sound),  ST08  (San  Carlos  Bay),  and  ST  10 
(Caloosahatchee  River  mouth) 1*1 

5.36  The  comparisons  of  surface  residual  flow  and  salinity  fields  for  three  cases.   ...  152 

5.37  The  comparisons  of  bottom  residual  flow  and  salinity  fields  for  three  cases.    ...  153 

5.38  The  vertical-longitudinal  salinity  profiles  along  the  axis  of  the  Caloosahatchee  River 

during  wet  season  in  2000 l->9 


XI 


5.39  The  vertical-longitudinal  salinity  profiles  along  the  axis  of  the  Caloosahatchee  River 

during  dry  season  in  2000 160 

5.40  Time  histories  of  river  discharge  at  S-79  and  the  locations  of  1,  10,  and  20  ppt 

surface  salinity  along  the  Caloosahatchee  River  during  2000  simulation  period.161 

5.41  Time  histories  of  river  discharge  at  S-79  and  the  1  ppt  salinity  location  along 

Caloosahatchee  River 162 

5.42  The  1-day  averaged  20  ppt  surface  salinity  location  and  30-day  averaged  10  ppt 

surface  salinity  location  during  2000  simulation  period 163 

5.43  The  locations  of  10  ppt  surface  salinity  due  to  river  discharge  rate  at  S-79  during 

2000  baseline  simulation 164 

5.44  The  relationship  between  location  of  specific  salinity  value  vs.  river  discharge  at 

S-79 165 

5  45  The  relationship  between  salinity  at  Fort  Myers  station  vs.  river  discharge  at  S-79. 

166 

6.1  Tidal  forcing  and  river  discharges  for  1996  simulations  of  Charlotte  Harbor 169 

6.2  Wind  velocity  for  1996  simulations  of  Charlotte  Harbor 170 

6.3  Air  temperature  for  1996  simulations  of  Charlotte  Harbor 171 

6.4  Locations  of  1996  water  quality  measurement  stations  operated  by  EPA   176 

6.5  Locations  of  2000  water  quality  measurement  stations  operated  by  SFWMD  and 

SWFWMD 177 

6.6  Light  intensity  at  water  surface  for  1996  and  2000  simulations 178 

6.7  Segments  for  Charlotte  Harbor  estuarine  system  179 

6.8  The  scatter  plots  for  water  quality  constituents  during  calibration  period  189 

6.9  Temporal  water  quality  variations  at  CH002  station  in  1996 194 

6.10  Temporal  water  quality  variations  at  CH004  station  in  1996 195 

6.1 1  Temporal  water  quality  variations  at  CH005  station  in  1996 196 

6.12  Temporal  water  quality  variations  at  CH006  station  in  1996 197 

xii 


6.13  Temporal  water  quality  variations  at  CH007  station  in  1996 198 

6.14  Temporal  water  quality  variations  at  CH09B  station  in  1996 199 

6.15  Temporal  water  quality  variations  at  CH009  station  in  1996 200 

6.16  Temporal  water  quality  variations  at  CH010  station  in  1996 201 

6.17  Temporal  water  quality  variations  at  HB002  station  in  1996 202 

6.18  Temporal  water  quality  variations  at  HB006  station  in  1996 203 

6.19  Temporal  water  quality  variations  at  HB007  station  in  1996 204 

6.20  Temporal  water  quality  variations  at  CH013  station  in  1996 205 

6.21  Simulated  dissolved  oxygen  concentration  in  Charlotte  Harbor  estuarine  system  for 

August  21,  1996 206 

6.22  Simulated  chlorophyll  a  concentration  in  Charlotte  Harbor  estuarine  system  for 

August  21,  1996 207 

6.23  Simulated  dissolved  ammonium  nitrogen  concentration  in  Charlotte  Harbor 

estuarine  system  for  August  21,  1996 208 

6.24  Simulated  soluble  organic  nitrogen  concentration  in  Charlotte  Harbor  estuarine 

system  for  August  21,  1996 209 

6.25  Simulated  soluble  reactive  phosphorous  concentration  in  Charlotte  Harbor  estuarine 

system  for  August  21,  1996 210 

6.26  Simulated  soluble  organic  phosphorous  concentration  in  Charlotte  Harbor  estuarine 

system  for  August  21,  1996 211 

6.27  The  scatter  plots  for  water  quality  constituents  in  2000 213 

6.28  Temporal  water  quality  variations  at  CH002  station  in  2000 218 

6.29  Temporal  water  quality  variations  at  CH004  station  in  2000 219 

6.30  Temporal  water  quality  variations  at  CH005  station  in  2000 220 

6.31  Temporal  water  quality  variations  at  CH006  station  in  2000 221 


xm 


6.32  Temporal  water  quality  variations  at  CH007  station  in  2000 222 

6.33  Temporal  water  quality  variations  at  CH09B  station  in  2000 223 

6.34  Temporal  water  quality  variations  at  CH009  station  in  2000 224 

6.35  Temporal  water  quality  variations  at  CH010  station  in  2000 225 

6.36  Temporal  water  quality  variations  at  CES02  station  in  2000 226 

6.37  Temporal  water  quality  variations  at  CES03  station  in  2000 227 

6.38  Temporal  water  quality  variations  at  CES08  station  in  2000 228 

6.39  Temporal  water  quality  variations  at  CHOI 3  station  in  2000 229 

6.40  The  comparison  between  simulated  dissolved  oxygen  concentration  and  the 

possible  causes  hypoxia:  river  discharge,  salinity,  temperature,  and  re-aeration  and 
SOD  fluxes  at  CH006  water  quality  measured  station 230 

6.41  Simulated  longitudinal-vertical  salinity  and  dissolved  oxygen  concentration  along 

the  Peace  River  at  1  pm  on  June  18  (Julian  Day  170),  2000 231 

6.42  Simulated  longitudinal-vertical  salinity  and  dissolved  oxygen  concentration  along 

the  Peace  River  at  1  pm  on  October  6  (Julian  Day  280),  2000 232 

6.43  Simulated  near-surface  chlorophyll  a  concentration  in  Charlotte  Harbor  estuarine 

system  for  February  9,  May  9,  August  7,  November  5,  2000 233 

6.44  Simulated  near-bottom  dissolved  oxygen  concentration  in  Charlotte  Harbor 

estuarine  system  for  February  9,  May  9.  August  7,  November  5,  2000 234 

6.45  Comparisons  of  simulated  surface  chlorophyll  a  concentration  for  three  cases  at 

three  selected  stations:  ST05  (Pine  Island  Sound),  ST08  (San  Carlos  Bay),  and 
ST10  (Caloosahatchee  River  mouth) 237 

6.46  Comparisons  of  simulated  surface  chlorophyll  a  concentration  fields  in  San  Carlos 

Bay  after  90  days  of  simulation  for  three  cases 238 

6.47  Comparisons  of  simulated  surface  dissolved  ammonium  nitrogen  (NH4) 

concentration  fields  in  San  Carlos  Bay  after  90  days  of  simulation  for  three  cases. 
239 

6.48  The  water  quality  species  at  CH004  water  quality  measured  station  before  and  after 

100  %  nitrogen  load  reduction  from  Peace  River 247 

xiv 


6.49  The  water  quality  species  at  CH006  water  quality  measured  station  before  and  after 

100  %  nitrogen  load  reduction  from  Peace  River 248 

6.50  The  water  quality  species  at  CH004  water  quality  measured  station  before  and  after 

100  %  phosphorous  load  reduction  from  Peace  River 249 

6.51  The  water  quality  species  at  CH006  water  quality  measured  station  before  and  after 

100  %  phosphorous  load  reduction  from  Peace  River 250 

6.52  The  water  quality  species  at  CES02  water  quality  measured  station  before  and  after 

100  %  nitrogen  load  reduction  from  Caloosahatchee  River 251 

6.53  The  water  quality  species  at  CES08  water  quality  measured  station  before  and  after 

100  %  nitrogen  load  reduction  from  Caloosahatchee  River 252 

6.54  The  water  quality  species  at  CES02  water  quality  measured  station  before  and  after 

100  %  phosphorous  load  reduction  from  Caloosahatchee  River 253 

6.55  The  water  quality  species  at  CES08  water  quality  measured  station  before  and  after 

100  %  phosphorous  load  reduction  from  Caloosahatchee  River 254 

6.56  Dissolved  oxygen  and  Chlorophyll  a  concentrations  at  CH006  water  quality 

measured  station  before  and  after  100  %  organic  matter  load  reduction  from  Peace 
River  using  DiToro's  sediment  flux  model 255 

6.57  Dissolved  oxygen  concentrations  at  CH004  and  CH006  water  quality  measured 

stations  before  and  after  50  %  SOD  reduction  and  75%  SOD  reduction 256 

6.58  The  comparison  of  hypoxia  area  at  Upper  Charlotte  Harbor  according  to  varying 

SOD  constant  rate  at  20°C 257 

A.l  Flow  chart  for  main  CH3D  program 264 

A.2  Flow  chart  for  driving  subroutine  for  the  time  stepping  of  the  solution 266 

A. 3  Flow  chart  for  initializing  sediment  transport  model 271 

A.4  Flow  chart  for  main  sediment  transport  model 272 

A.5  Flow  chart  for  initializing  water  quality  model  277 

A.6  Flow  chart  for  main  water  quality  model 275 


xv 


A.7  Flow  chart  for  computing  temperature  and  light  attenuation  functions  for  water 

quality  model 277 

D.l  The  vertical  one-dimensional  z-grid   292 

E.  1  Nitrogen  cycle 296 

E.2  Phosphorous  cycle 302 

G.l  Parallel  speedup  gained  in  performing  the  simulation  on  the  SGI  Origin  platform.312 


xvi 


Abstract  of  Dissertation  Presented  to  the  Graduate  School 

of  the  University  of  Florida  in  Partial  Fulfillment  of  the 

Requirements  for  the  Degree  of  Doctor  of  Philosophy 

MODELING  THE  CIRCULATION  AND  WATER  QUALITY 
IN  CHARLOTTE  HARBOR  ESTUARINE  SYSTEM,  FLORIDA 

By 

Kijin  Park 
August  2004 

Chair:  Y.  Peter  Sheng 

Department:  Civil  and  Coastal  Engineering 

This  study  aims  to  develop  an  enhanced  version  of  a  three-dimensional  curvilinear- 
grid  modeling  system,  CH3D-IMS,  which  include  a  3-D  hydrodynamics  model,  a  3-D 
sediment  transport  model,  and  a  3-D  water  quality  model,  to  simulate  circulation  and  water 
quality  of  the  Charlotte  Harbor  Estuarine  System  and  to  provide  quantitative  assessment  of 
various  management  practices. 

In  the  past  decade,  the  upper  Charlotte  Harbor  system  has  been  suffering  summer 
hypoxia  in  bottom  water.  Field  study  indicated  that  hypoxia  in  the  upper  Charlotte  Harbor 
is  related  to  a  strong  stratification  caused  by  high  freshwater  flows  and  dissolved  oxygen 
fluxes  at  the  air-sea  and  sediment-water  interfaces.  To  simulate  the  hypoxia  event,  models 
of  oxygen  balance  and  oxygen  fluxes  at  the  air-sea  and  sediment-water  interfaces  in  previous 
versions  of  CH3D-IMS  are  enhanced.  The  three  dimensional  temperature  model  and 
physics-based  light  model  in  CH3D-IMS  are  also  enhanced  to  enable  better  understanding 


xvn 


of  the  temporal  and  spatial  variations  of  temperature  and  light  and  their  effect  on  water 
quality  processes. 

The  hydrodynamics  component  of  the  integrated  model  CH3D-IMS  for  Charlotte 
Harbor  has  been  successfully  calibrated,  using  hydrodynamic  data  gathered  by  the  National 
Oceanic  and  Atmospheric  Administration  (NOAA)  and  the  United  States  Geological  Survey 
(USGS)  in  1986  and  2000.  This  calibrated  model  was  applied  to  assess  the  impact  of  the 
removal  of  the  Sanibel  Causeway  and  the  IntraCoastal  Waterway  on  the  circulation  in  the 
San  Carlos  Bay  area  and  to  provide  a  quantitative  evaluation  of  the  minimum  flow  and  level 
(MFL)  for  Caloosahatchee  River.  The  results  show  that  the  hydrologic  alterations  would  not 
noticeably  affect  the  circulation,  salinity,  and  water  quality  in  San  Carlos  Bay  except  near  the 
mouth  of  Caloosahatchee  River.  The  minimum  flow  required  to  produce  a  salinity  of  no 
more  than  10  ppt  at  Fort  Myers  is  about  18  m3/s. 

The  water  quality  model  of  the  CH3D-MS  was  calibrated  with  the  systematic 
calibration  procedure  and  validated  using  hydrodynamic,  sediment,  and  water  quality  data 
provided  by  the  USGS  and  the  United  States  Environmental  Protection  Agency  (USEPA) 
in  1996,  and  by  the  USGS,  the  South  Florida  Water  Management  District  (SFWMD)  and  the 
Southwest  Florida  Water  Management  District  (SWFWMD)  in  2000.  This  validated  model 
was  used  to  examine  the  temporal  and  spatial  dynamics  of  factors  which  can  affect  hypoxia 
in  upper  Charlotte  Harbor,  such  as  freshwater  inflow,  tidal  variation,  sediment  oxygen 
demand  (SOD),  water  column  oxygen  consumption,  and  dissolved  oxygen  (DO)  re-aeration. 
The  model  results  suggest  that  hypoxia  in  the  upper  Charlotte  Harbor  System  is  primarily 
caused  by  a  combination  of  vertical  salinity  stratification  and  SOD,  while  water  quality  also 
affected  by  the  oxygen  re-aeration  and  water  column  oxygen  consumption.  This  model  was 


xvin 


applied  to  assess  the  effects  of  hydrologic  alterations  and  to  provide  a  preliminary  evaluation 
of  pollutant  load  reduction  goal  (PLRG).  Due  to  lack  of  detailed  data  and  insufficient 
understanding  on  the  causes  of  SOD  and  how  SOD  is  related  to  the  loading,  the  present 
model  cannot  simulate  the  complete  effect  of  nutrient  load  reduction  on  the  DO 
concentration  in  the  estuary.  Further  research  and  more  complete  data  are  needed 

A  systematic  calibration  procedure  has  been  developed  for  a  more  efficient  and  more 
objective  calibration  of  the  water  quality  model.  A  consistent  framework  for  systematic 
calibration  is  formulated,  which  include  in  the  following  steps:  model  parameterization, 
selection  of  calibration  parameters,  and  formulation  of  calibration  criteria. 


xix 


CHAPTER  1 
INTRODUCTION 

Charlotte  Harbor  (Figure  1 . 1 )  is  a  shallow  estuarine  system  in  southwest  Florida.  The 

estuary  receives  freshwater  from  the  Caloosahatchee,  Peace,  and  Myakka  Rivers;  is 

connected  to  the  Gulf  of  Mexico  through  the  Boca  Grande  Pass,  Gasparilla  Pass,  Captiva 

Pass,  Blind  Pass,  and  San  Carlos  Bay;  and  provides  water  resources  for  several  counties  in 

Southwest  Florida.  Charlotte  Harbor  is  dominated  by  rivers  that  flow  into  the  coastal  area. 

While  most  estuaries  in  Southwest  Florida  are  influenced  by  the  Gulf  of  Mexico,  the 

characteristics  of  the  Charlotte  Harbor  system  are  strongly  influenced  by  large  rivers  such  as 

the  Caloosahatchee  and  Peace  Rivers.  Large  fluctuations  of  river  flow  between  wet  and  dry 

seasons  strongly  affect  the  salinity  and  water  characteristics  in  Charlotte  Harbor  (Estevez, 

1998). 

Population  growth  and  development  in  the  surrounding  areas  during  the  past  few 
decades  have  led  to  concerns  over  human  impacts  on  the  quality  of  the  estuarine  system. 
Industrial  and  agriculture  development  also  increase  environment  pollution.  Growth  and 
development  will  cause  an  increased  demand  for  fresh  water  and  a  corresponding  increase 
in  urban,  agricultural,  and  industrial  waste.  The  inflow  of  freshwater  is  essential  to  the 
integrity  and  health  of  the  estuarine  system.  Increased  freshwater  withdrawal  or  diversion, 
or  increased  wastewater  discharges  to  the  rivers  and  streams  that  flow  into  the  estuary  will 
create  environmental  stress  in  the  estuary  (McPherson  et  al.,  1996). 


1 


Myakka  River 


%>     Deer  Prairie  Creek 

\  Dirt  Cl^i  tr 


Big  Slough  Canal 


h&  J      Sarasota  Co. 

Nortlmortf*Z- •--- 

■y  Char 


Peace  River 


!\  V  El  Jpbean 


i        j«     De_SotoC_o. 

clia7iotteCo.*W'~ 

-.  11 


\^Gaspar1lla\   "«*»    i 
w^  Sound      ~  Charlotte  jr 

Gasparilla  \h£ ,.)     M  Harbor  It- 
Pass  -^  »iw/''fi  £ 


V^Lower    X 

-rravfa  Charlotte  HarborV  / 
Boca  Grande^  Bokeelia   y. 

Pass         ,  -.  ^^M^f% 

CayoCostit.':  S  ^ 


S79 
Caloosahatchee  River.  fJmQj        <T'" 


Charlotte  Co. 
Lee  Co. 


'u^  &         ^ -\  Matlacha  Pass  J- *  ~X 
Captiva    'Pin_,-^i    \      i*„  ■;    --.  Fort  Myers 

Pjjqq      ^-     I        II  IC  -i  .-  j,  ■■■■■       i,  < _, 

rass  Island  ..    v    &M     r„„„    tJP    ;' 


■Island  .\  \j?f>t  Cape  jgP  / 
v  Sound  i  #  a*  C'or«/.  >f  ,> 
RedfishV  -T  -.v       aia^.^vv 


'-■:,  Sound-'^,  %"'  £*     Cora/  ^"|f     ,> 
^  Cf    V':\-.       StlJ^    /^Whiskey Creek 

f,    PineSf  1^-^r^^ 
Captiva  ,-,  Islanq  Ajlffr  ._  *piVtr..  _  .        f 


Pass 


Blind 
Pass- 


o 


$iell  PoiiU    j 

-■ .-  „-* 

Sanibel/       ./''     Causeway  '%r^t 

Island       San  Carlos  Fort  M      s^fc  ^ 
Bay  &at./i        <$  j 


>P- •■'■>->  M 

Estero  Bay  *5r       Imperial  River       ■ * 

:%"~         A 


EXPLANATION 

— —  Subarea  Divide 


1 
»  Naples 


Kilometers 


Lb  ^  ^        e 

ho  0  10 


20 


_^__ i . 

Figure  1.1  Charlotte  Harbor  estuarine  system  and  its  subarea  boundaries 


Water  quality  in  the  southern  Charlotte  Harbor  estuarine  system  (the  Caloosahatchee 
Estuary,  San  Carlos  Bay,  and  Pine  Island  Sound)  appear  to  be  influenced  by  freshwater 
discharge  from  the  Franklin  Lock  and  Dam,  also  known  as  United  States  Army  Corps  of 
Engineers  Structure  Number  79  (S-79).  Waste  load  allocation  studies  conducted  by  the 
Florida  Department  of  Environmental  Regulation  (Degrove,  198 1 ;  Degrove,  and  Nearhoof, 
1987;  Baker,  1990)  concluded  that  the  Caloosahatchee  Estuary  had  reached  its  nutrient 
loading  limits  as  indicated  by  elevated  chlorophyll-a  and  depressed  dissolved  oxygen  (DO) 
levels.  Similarly,  McPherson  and  Miller  (1990)  concluded  that  increased  nitrogen  loading 
would  result  in  undesirable  increases  in  phytoplankton  and  benthic  algae.  In  order  to 
effectively  manage  the  loading  of  pollutants  such  as  nutrients  in  a  shallow  estuary  such  as 
Charlotte  Harbor,  and  control  eutrophication,  it  is  necessary  to  have  a  quantitative 
understanding  of  the  transport  and  transformation  processes  of  nutrients  in  system. 

Increased  nutrient  loading  from  tributaries,  the  atmosphere,  and  bottom  sediments 
have  been  known  to  cause  eutrophication  in  estuaries.  Eutrophication  varies  both  spatially 
and  temporally.  Eutrophication  does  not  necessarily  occur  in  regions  with  high  nutrient 
loading  because  the  circulation  and  sediment  dynamics  affect  nutrient  fate  and  transport.  In 
the  Upper  Charlotte  Harbor  where  a  pycnocline  develops  frequently  during  high  river  flow 
conditions,  hypoxia  usually  develops  within  two  days  of  the  formation  of  the  pycnocline, 
when  coupled  with  significant  sediment  oxygen  demand  (SOD)  (CDM,  1998).  In  fact, 
hypoxia  can  occasionally  extend  all  the  way  to  the  Boca  Grande  area.  This  strong  linkage 
between  hydrodynamics  and  water  quality  dynamics  suggests  the  importance  of  having  a 
comprehensive  understanding  of  hydrodynamics. 

Nutrient  dynamics  in  estuaries  are  not  only  determined  by  biological  and  chemical 


processes,  but  is  also  strongly  affected  by  weather  and  climate  (wind,  air  temperature,  etc.), 
hydrodynamics  (wave,  tide,  current,  turbulence  mixing,  etc.),  and  sediment  transport 
processes  (resuspension,  deposition,  flocculation,  etc.).  In  this  study,  integrated  modeling 
of  the  system  is  conducted  to  understand  the  complex  water  quality  processes,  which  are 
strongly  linked  to  hydrodynamics  and  sediment  processes.  The  primary  objective  of  this 
study  is  to  use  models  and  field  data  to  produce  a  detailed  characterization  of 
hydrodynamics,  sediment,  and  water  quality  dynamics  within  the  Charlotte  Harbor  estuarine 
system. 

The  Chesapeake  Bay  model  (Cerco  and  Cole,  1994)  and  Moreton  Bay  model 
(McEwan  and  Garbric,  1998)  are  coupled  hydrodynamics-water  quality  models.  These 
coupled  models  are  more  suitable  for  managing  nutrient  loads  and  predicting  eutrophication- 
related  problems  than  uncoupled  models.  Because  these  models  were  not  coupled  with  a 
dynamic  model  for  sediment  transport,  however,  they  could  not  accurately  consider 
sediment-process  effects  such  as  resuspension,  deposition,  flocculation,  and  settling  on 
nutrient  dynamics  in  estuaries.  Therefore,  these  loosely  coupled  models  cannot  account  for 
nutrient  release  by  sediments  in  episodic  events  (Chen  and  Sheng,  1994) 

Chen  and  Sheng  (1994)  developed  a  coupled  hydrodynamic-sediment-water  quality 
model  and  applied  it  to  Lake  Okeechobee.  Their  3-D  model  includes  a  hydrodynamic  model, 
a  sediment  transport  model,  and  a  water  quality  model  with  a  nitrogen  and  a  phosphorous 
cycle  on  a  rectangular  grid.  Measured  nitrogen  and  phosphorous  dynamics  during  episodic 
events  in  1993  and  monthly  sampling  events  in  1989  were  accurately  simulated  with 
absorption-desorption  reactions  and  the  exchange  of  nutrients  between  the  sediment  and 
water  column. 


While  Chen  and  Sheng  (1994)  developed  a  rectangular  grid  model,  Yassuda  and 
Sheng  (1996)  developed  an  integrated  model  in  curvilinear  grids  and  applied  it  to  Tampa 
Bay.  Their  water  quality  model,  based  on  the  Chen  and  Sheng  (1994)  model,  includes  a 
nitrogen  cycle  and  an  oxygen  cycle,  but  not  a  phosphorous  cycle.  Zooplankton  distribution 
is  allowed  to  influence  the  phytoplankton  dynamics.  Their  integrated  model  also  includes 
a  wave  model,  a  light  model,  and  a  seagrass  model.  Although  Yassuda  and  Sheng  (1996) 
simulated  the  observed  hypoxia  in  Tampa  Bay  during  1991,  their  model  did  not  include 
dynamic  fluxes  of  oxygen  at  the  air-sea  interface  (re-aeration)  and  the  sediment-water 
interface  (sediment  oxygen  demand).  Their  model  could  not  simulate  the  daily  fluctuation 
and  vertical  stratification  of  DO.  Although  the  phytoplankton  growth  rate  in  their  model  is 
controlled  by  temperature  and  light,  it  is  difficult  to  reproduce  the  vertical  distribution  of 
phytoplankton  because  the  temperature  is  not  simulated.  The  light  model  in  Yassuda  and 
Sheng  (1996)  has  not  been  sufficiently  validated  with  Tampa  Bay  data,  thus  limiting  the 
predictability  of  their  model  for  vertical  distribution  of  phytoplankton  and  dissolved  oxygen. 
Their  model  could  not  produce  the  low  dissolved  oxygen  phenomena  in  upper  Charlotte 
Harbor  because  there  are  no  dissolved  oxygen  fluxes  at  the  air-water  and  sediment-water 
interfaces,  such  as  reaeration  and  sediment  oxygen  demand  (SOD). 

Sheng  (2000)  developed  the  framework  for  an  integrated  modeling  system:  CH3D- 
EVIS  (http://ch3d.coastal.ufl.edu).  The  integrated  modeling  system  (Table  1.1)  is  based  on 
the  curvilinear-grid  hydrodynamic  model  CH3D  (Sheng,  1987, 1989;  Sheng  et  al.,  2002)  and 
also  includes  a  wave  model,  a  sediment  transport  model  (Sheng  et  al.,  2002a),  a  water  quality 
model,  (Sheng  et  al.,  2001b),  a  light  attenuation  model  (Sheng  et  al.,  2002c,  Christian  and 
Sheng,  2003),  and  a  seagrass  model  (Sheng  et  al.,  2002d).  Sheng  et  al.  (2002)  applied  the 


CH3D-IMS  to  the  Indian  River  Lagoon  (IRL),  Florida.  The  water  quality  model  (Sheng  et 
al.,  2001b)  includes  a  nitrogen  cycle,  a  phosphorous  cycle,  a  phytoplankton  cycle,  and  a  DO 
cycle.  The  CH3D-IMS  was  validated  with  comprehensive  data  collected  from  the  IRL. 
Table  1.1  Component  Models  of  the  CH3D-IMS 


Component  Model  Model  Name 


Hydrodynamic  Model  CH3D 

Flow  Model 
Salinity  Transport  Model 

Wave  Model  SMB 

Sediment  Transport  Model  CH3D-SED3D 

Water  Quality  Model  CH3D-WQ3D 

Dissolved  Oxygen  Model 

Phytoplankton  and  zooplankton 
Model 

Nitrogen  Model 

Phosphorous  Model 

Light  Attenuation  Model  CH3D-LA 

Seagrass  Model  CH3D-SAV 


The  CH3D-IMS  was  able  to  successfully  simulate  the  observed  circulation,  wave, 
sediment  transport,  water  quality,  light  attenuation,  and  seagrass  biomass  in  IRL  during  1998. 
However,  several  aspects  of  the  model  were  identified  for  further  improvement.  For 
example,  although  a  temperature  model  was  developed  during  the  later  phase  of  the  IRL 
study,  it  was  not  sufficiently  validated  with  data.  The  water  quality  model,  while  capable  of 
simulating  the  annual  variation  in  DO,  could  not  simulate  the  diurnal  DO  variation. 

In  this  study,  the  CH3D-IMS  is  enhanced  and  used  to  study  the  circulation  and  water 
quality  dynamics  in  the  Charlotte  Harbor  estuarine  system.  The  validated  model  will  be  used 
to  determine  minimum  flow  and  level  and  pollutant  load  reduction  goal  by  management 
agencies.     Specially,  the  CH3D-IMS  used  for  this  study  includes  the  simulation  of 


temperature  in  addition  to  the  flow  and  salinity  simulations.  An  air-sea  heat  flux  model  is 
implemented  along  with  the  temperature  equation.  The  water  quality  model  is  enhanced  to 
include  a  re-aeration  model  and  a  sediment  oxygen  demand  (SOD)  model,  such  that  hypoxia 
and  daily  fluctuation  and  vertical  stratification  of  DO  can  be  simulated.  Moreover,  the  same 
basic  light  attenuation  model  developed  for  the  IRL  is  used  for  this  study. 

Although  the  factors  that  cause  hypoxia  have  not  been  positively  identified,  it  is 
generally  accepted  that  hypoxia  is  related  to  strong  stratification  caused  by  high  freshwater 
inflow.  The  modified  CH3D-IMS  is  used  in  this  study  to  examine  the  major  factors  of  the 
observed  hypoxia  in  2000,  including  freshwater  inflow,  tidal  variation,  SOD,  water  column 
oxygen  consumption,  and  DO  reaeration. 
This  study  includes  the  following  objectives: 

•  To  validate  the  CH3D-IMS  with  extensive  hydrodynamic  and  water  quality  data  from 
Charlotte  Harbor, 

To  improve  the  simulation  of  dissolved  oxygen  processes,  including  the  observed 
daily  dissolved  oxygen  fluctuation,  and  hypoxia  in  bottom  water  during  high  flow 
event  with  vertical  stratification, 

•  To  validate  the  ability  of  temperature  model  and  heat  flux  model  to  simulate  seasonal 
and  spatial  temperature  distribution  in  Charlotte  Harbor, 

•  To  apply  the  physics-based  light  attenuation  model  to  improve  Charlotte  Harbor 
simulations,  and 

To  demonstrate  the  feasibility  of  the  validated  modeling  system  for  simulating  the 
effects  of  various  anthropogenic  impacts  on  the  Charlotte  Harbor  estuarine  system. 

More  specifically,  the  objectives  of  my  research  are  to: 

•  Reproduce  observed  circulation  and  transport  dynamics  in  1986  and  2000, 

•  Reproduce  observed  water  quality  dynamics  in  1996  and  2000, 

•  Investigate  the  effect  of  the  Sanibel  causeway  islands  and  IntraCoastal  Waterway  on 
the  flow,  salinity,  and  water  quality  distribution  in  San  Carlos  Bay  and  Pine  Island 


Sound, 

Reproduce  the  main  factors  for  bottom  water  hypoxia  and  habitat  loss  in  Charlotte 
Harbor  estuarine  system, 

•  Provide  a  preliminary  determination  of  pollutant  load  reduction  goal  (PLRG)  and 
minimum  flow  and  levels  (MFL), 

Perform  sensitivity  tests  to  quantify  the  influence  of  each  model  parameter  and  to 
determine  the  most  sensitive  parameters  and  most  plausible  parameter  set,  and 

•  Establish  a  systematic  calibration  procedure. 

To  achieve  these  goals,  the  characteristics  of  Charlotte  Harbor  estuarine  system  are 
presented  in  Chapter  2.    Chapter  3  discusses  the  general  hydrodynamic  circulation  and 
transport  of  temperature,  salinity,  and  sediment  in  the  estuarine  system.  This  chapter  will 
include  initial  and  boundary  conditions  for  the  hydrodynamic  and  transport  models.  As  an 
air/sea  interface  boundary  condition,  a  heat  flux  model  is  also  presented  in  this  chapter. 
Chapter  4  describes  the  water  quality  processes  and  models  of  phytoplankton,  zooplankton, 
nitrogen,  phosphorous,  dissolved  oxygen,  and  light  attenuation  processes  in  estuarine  system. 
The  effects  of  temperature  and  light  intensity,  numerical  solution  technique,  and  model 
parameters  and  calibration  procedure  are  also  discussed  in  this  chapter.  Chapter  5  presents 
hydrodynamic  field  data  and  model  simulations  of  hydrodynamics  in  Charlotte  Harbor 
estuarine  system.  These  model  simulations  include  model  calibration  and  verification  of 
1986  and  2000  data,  simulation  of  the  effect  of  causeway  islands  and  navigation  channel,  and 
determination  of  minimum  flow  and  levels  (MFL).  Chapter  6  presents  water  quality  and 
sediment  field  data  and  model  simulations  of  circulation,  sediment  transport  and  water 
quality  processes  in  the  study  area.  These  model  simulations  include  calibration  and 
verification  of  1996  and  2000  data  and  determination  of  pollutant  load  reduction  goal 
(PLRG).  Discussions  and  conclusions  are  presented  in  Chapter  7. 


CHAPTER  2 
CHARLOTTE  HARBOR  CHARACTERIZATION 

Charlotte  Harbor,  a  coastal-plain  estuarine  system  is  one  of  the  largest  estuarine 
systems  on  the  southwest  florida  coast  and  is  an  important  part  of  the  Gulf  of  Mexico 
watershed.  It  is  located  on  the  southwest  corner  of  the  Florida  peninsula,  between  26°  20' 
and  27°  10'N  and  81°  40'  and  82°  30'  W.  As  shown  in  Figure  1.1,  the  Charlotte  Harbor 
estuarine  system  is  sub-divided  into  Upper  Charlotte  Harbor,  Lower  Charlotte  Harbor,  Pine 
Island  Sound,  Matlacha  Pass,  San  Carlos  Bay,  Gasparilla  Bay,  Peace  River,  Myakka  River 
and  Caloosahatchee  River.  The  drainage  area  (Figure  2.1)  consists  of  the  Peace,  Myakka, 
and  Caloosahatchee  River  Watersheds  and  the  coastal  area  and  islands  that  drain  directly  into 
the  harbor.  The  estuary  has  a  surface  area  of  648  km2,  a  drainage  area  of  a  more  than  10000 
km2,  and  a  average  depth  of  2. 1  m.  The  upper  harbor  has  an  average  depth  of  2.6  m,  and  the 
lower  harbor  has  an  average  depth  of  1.6  m  (Stoker,  1986).  The  estuary  is  separated  from 
the  Gulf  of  Mexico  by  barrier  islands  and  is  connected  to  the  Gulf  through  two  major  inlets: 
Boca  Grande  and  San  Carlos;  and  through  several  smaller  passes:  Gasparilla,  Captiva, 
Redfish  (McPherson  et  al.,  1996). 

In  1995,  the  Charlotte  Harbor  National  Estuary  Program  (CHNEP)  was  established 
by  U.S.  Environmental  Protection  Agency  (USEPA)  and  the  State  of  Florida.  In  its  planning 
documents,  CHNEP  identified  three  draft  major  problems:  (1)  hydrologic  alterations,  (2) 
nutrient  enrichment,  and  (3)  habitat  loss.  In  order  to  address  these  problems,  it  is  useful  to 


10 
have  a  comprehensive  understanding  of  the  climate,  hydrodynamics,  sediment  properties,  and 
biological  and  chemical  characteristics  of  the  system. 


27°45' 


27°00'  - 


Figure  2.1  Drainage  basins  of  Charlotte  Harbor  estuarine  system  (Stoker,  1992) 


11 

2.1  Climate 

The  climate  of  the  Charlotte  Harbor  estuarine  system  is  subtropical  and  humid.  The 
mean  annually  average  temperature  is  72  °F,  with  low  mean  of  60°F  in  December  and 
January  and  a  high  mean  of  80  °F  during  the  summer  (McPherson  et  al.,  1996). 

Annual  rainfall  averages  132  cm,  of  which  more  than  half  occurs  from  June  through 
September,  during  local  thundershowers  and  squalls.  Rain  during  the  fall,  winter,  and  spring 
seasons  is  usually  the  result  of  large  frontal  systems  and  tends  to  be  more  broadly  distributed 
than  rain  associated  with  local  thunderstorms  and  squalls.  The  period  from  October  through 
February  is  characteristically  dry,  with  November  usually  being  the  driest  month.  The 
months  of  April  and  May  also  are  characteristically  dry.  Low  rainfall  in  April  and  May 
coincides  with  high  evaporation  and  generally  results  in  the  lowest  streamflow,  lake  stage, 
and  ground-water  levels  of  the  year  (Hammett,  1990). 

The  annual  average  wind  speed  is  3.9  m/s  from  the  east.  Four  typical  seasonal  wind- 
field  patterns  are  shown  in  Figure  2.2.  In  the  Winter  months,  the  easterly  trade  winds 
dominate  the  region  south  of  latitude  27°  N,  while  the  westerlies  dominate  the  area  north  of 
latitude  29°N.  Spring  and  Summer  generally  exhibit  more  southerly  winds,  and  Fall  is 
characterized  by  easterly  or  northeasterly  winds.  Wind  speed  can  exceed  10  m/s  during  the 
passage  of  Winter  storms  or  during  Summer  squalls,  hurricanes  and  tornadoes  (Wolfe  and 
Drew,  1990). 

2.2  Hydrodynamics 

The  hydrodynamic  behavior  of  the  Charlotte  Harbor  estuarine  system  is  affected  by 
physical  characteristics  (bathymetry  and  geometry)  as  well  as  weather,  climate,  and 
oceanographic  and  hydrologic  characteristics,  such  as  wind,  atmosphere  heating  and  cooling, 


12 
evaporation  and  precipitation,  tidal-stage  oscillations,  discharge  through  tidal  inlets,  tidal 
velocity,  and  freshwater  inflow. 


Figure  2.2  Seasonal  wind  pattern  in  Florida  (Echternacht,  1975) 


13 
2.2.1  Tidal  Stage,  Discharge  at  Inlet,  and  Tidal  Circulation 

Tides  along  the  Gulf  coast  of  West-Central  Florida  in  the  vicinity  of  Charlotte  Harbor 
have  a  range  of  30  to  140  cm  and  are  of  the  mixed  type  with  both  diurnal  and  semidiurnal 
characteristics  (Goodwin  and  Michaelis,  1976).  Spring  tides,  which  have  the  largest  range, 
sometimes  have  only  a  diurnal  fluctuation,  whereas  neap  tides,  which  have  the  smallest 
range,  approach  semidiurnal  conditions  of  two  nearly  equal  high  and  low  water  levels  per 
day. 

The  boundary  between  the  Charlotte  Harbor  estuarine  system  and  the  Gulf  of  Mexico 
extends  about  40  mi  from  Gasparilla  Pass  on  the  north  to  San  Carlos  Pass  on  the  south. 
Tidal  characteristics  in  the  Gulf  of  Mexico  are  nearly  uniform  from  Gasparilla  Island  to  the 
western  face  of  Sanibel  Island,  but  are  of  a  larger  range  off  the  southern  shore  of  Sanibel 
Island  (Goodwin,  1996). 

Circulation  in  the  system  is  primarily  driven  by  Gulf  tides,  entering  the  system 
through  San  Carlos  Pass,  Boca  Grande  Pass,  Captita  Pass,  Redfish  Pass,  Gasparilla  Pass  and 
Blind  Pass.  San  Carlos  Pass  has  a  maximum  depth  of  5.3  m  and  a  width  of  3.3  km;  Boca 
Grande  Pass  has  a  maximum  depth  of  a  18.3  m  and  a  width  of  1.28  km.  The  discharge 
through  Boca  Grande  Pass  is  about  twice  the  discharge  through  San  Carlos  Bay  and  three  to 
four  times  the  discharge  through  Captiva  and  Redfish  Passes  (Goodwin,  1996).  The 
geometric  narrowing  that  occurs  at  passes  focuses  tidal  energy,  resulting  in  high  velocities 
associated  with  the  large  volume  of  water  moving  through  the  pass.  This  tidal  energy  is 
dispersed  inside  the  harbor  and  influences  the  harbor  and  tidal  rivers  as  much  as  40  to  44  km 
upstream  from  the  Peace  River  mouth.  There  is  about  a  2-hour  lag  between  tide  phases  at 
Boca  Grande  and  tide  phase  in  the  upper  harbor  near  the  mouth  of  the  Peace  River  (Stoker, 


14 
1992).  Tide  and  wind  keep  the  water  column  well  mixed  in  the  southern  part  of  the  estuary. 
In  the  northern  part  of  the  estuary,  vertical  stratification  can  develop  during  moderate  to  high 
fresh  water  inflows  and  can  persist  for  weeks  after  a  high  freshwater  inflow  event  (Sheng, 
1998). 

2.2.2  Freshwater  Inflow 

The  majority  of  the  freshwater  that  enters  Charlotte  Harbor  come  from  the  Myakka 
River,  Peace  River  and  Caloosahatchee  River.  Average  flows  are  17.8  m3/s,  56.9m3/s  and 
56.7  m3/s,  respectively.  Flows  in  the  Myakka  and  Peace  Rivers  are  largely  unregulated, 
while  flow  in  the  Caloosahatchee  is  controlled  by  operation  of  the  Franklin  Lock  about  43 
km  upstream  from  the  mouth.  Discharge  in  the  Peace  and  Myakka  Rivers  tend  to  peak  in 
August  and  September  when  rainfall  totals  are  generally  the  greatest.  Discharges  are  usually 
lowest  in  April  and  May  (Stoker,  1992).  The  Caloosahatchee  River  discharge  does  not 
always  correspond  to  rainfall  patterns  in  the  basin,  since  it  is  controlled  by  S-79. 

Analyses  of  long-term  streamflow  trends  in  the  Charlotte  Harbor  area  have  indicated 
statistically  significant  decreases  in  streamflow  at  several  gages  in  the  Peace  River  basin 
from  1931  to  1984  (Hammett,  1990).  The  long-term  decrease  in  streamflow  of  the  Peace 
River  is  probably  related  to  the  increased  use  of  ground  water  and  subsequent  decline  of  the 
potentiometric  surface  of  the  upper  Florida  aquafer  (Hammett,  1990).  A  sustained 
significant  reduction  in  streamflow  could  result  in  an  increase  of  salinity  in  Upper  Charlotte 
Harbor,  possibly  approaching  the  Gulf  of  Mexico  salinity  (Sheng,  1998).  The  impact  of 
freshwater  reduction  is  of  major  ecological  and  economic  significance. 

2.2.3  Salinity 

As  in  any  other  typical  estuarine  system,  Charlotte  Harbor  generally  exhibits 


15 
significant  horizontal  gradients  in  salinity.  The  higher  salinity  values  in  the  adjacent  Gulf 
of  Mexico  fluctuate  around  36  ppt,  whereas  the  lower  salinity  levels  nearly  zero  in  wet 
season  occur  near  the  mouth  of  creeks  and  rivers.  Seasonal  changes  in  salinity  occur 
primarily  in  response  to  changes  in  freshwater  inflow  from  the  Peace,  Myakka,  and 
Caloosahatchee  River  basins.  Other  sources  of  freshwater,  including  direct  rainfall,  runoff 
form  coastal  areas,  ground-water  seepage,  and  domestic  influence,  have  smaller  and  usually 
more  local  effects  on  salinity  in  the  estuary  (McPherson  et  al.,  1996). 

Stoker  (1992)  described  salinity  characteristics  in  the  system  based  on  data  collected 
from  June  1982  to  May  1987.  Salinity  generally  is  the  lowest  during  the  wet  season  between 
July  and  September,  and  is  the  highest  during  the  dry  season  from  January  through  March. 
Salinity  also  varies  daily  in  response  to  tidal  fluctuation.  Peak  salinity  is  near  the  flood-tide 
stage,  and  lowest  salinity  is  near  the  ebb-tide  stage. 

Due  to  the  shallow  depth  and  because  of  significant  vertical  mixing,  salinity  is 
generally  not  stratifed  in  the  southern  part  of  the  estuary.  Vertical  salinity  stratification  in 
the  Upper  Charlotte  Harbor  is  a  common  seasonal  occurrence  (Environmental  Quality 
Laboratory,  Inc.,  1979).  In  high  river  inflow  events,  a  stable  vertical  salinity  gradient  is 
created  which  suppresses  vertical  mixing  unless  there  are  sufficient  mixing  by  wind  or  tide. 

2.3  Water  Quality 

Water  quality  refers  to  the  condition  of  water  (e.g.,  dissolved  oxygen  concentration, 
chlorophyll  concentration,  light  attenuation,  etc.)  relative  to  legal  standard,  social 
expectations  or  ecological  health.  Overall,  water  quality  in  the  Charlotte  Harbor  estuarine 
system  is  fair  or  good,  but  some  areas  have  poor  water  quality  or  declining  trends  (Estevez, 
1998).  The  water  quality  of  an  estuary  is  strongly  influenced  by  hydrodynamic  processes 


16 

(e.g.,  circulation  and  flushing),  and  chemical  and  biological  processes  (e.g.,  ammonification, 
mineralization,  decomposition,  algae  uptake,  excretion,  and  mortality,  etc.)  in  the  estuary  and 

basin. 

Color  levels  and  concentrations  of  nitrogen,  phosphorous  and  chlorophyll-a  in 
Charlotte  Harbor  exhibit  pronounced  salinity-related  gradients  which  extend  from  the  head 
of  the  estuary  to  its  mouth  (Morrison,  1997).  The  water  quality  in  the  Caloosahatchee  system 
is  more  degraded  than  the  water  quality  in  the  Myakka  or  Peace  systems.  Oxygen  depletion 
is  common  upstream  of  Franklin  Lock.  Nutrient  and  chlorophyll  levels  are  high,  and  algal 
blooms  occur  regularly  in  the  tidal  river  (Estevez,  1998). 
2.3.1  Nutrients 

Nutrient  availability  is  a  key  factor  in  the  regulation  of  primary  productivity  in 
estuarine  and  coastal  water  (Ketchum,  1967).  Increased  nutrient  loads  related  to  the  urban 
development  of  coastal  basins  have  been  implicated  in  estuarine  nutrient  enrichment, 
increased  phytoplankton  productivity,  and  incresed  phytoplankton  biomass  (Jaworski,  1981). 
and  in  declines  of  seagrass  communities  (Orth  and  Moore,  1983). 

The  distribution  of  nutrients  in  the  system  is  mainly  the  result  of  nutrient  input  from 
rivers,  freshwater  and  tidal  flushing,  and  recycling  processes  in  the  estuary  (McPherson  and 
Miller,  1990).  The  major  factor  that  influences  estuarine  nutrient  distribution  is  freshwater 
inflow  from  rivers,  which  contributes  substantial  nutrient  loads  and  flushes  nutrients 
seaward.  NOAA  estimates  that  Charlotte  Harbor  receives  about  2,500  tons  of  nitrogen  as 
total  Kjeldahl  nitrogen,  or  TKN,  and  1,000  tons  of  phosphorous  per  year.  Relative  to  its 
dimensions  and  flushing  characteristics,  phosphorous  loads  are  high,  signifying  a  nitrogen 
limited  system  (Estevez,  1986). 


17 
Molar  ratios  of  dissolved  inorganic  nitrogen  to  dissolved  inorganic  phosphorous  were 
below  the  Redfield  ratio  of  16,  averaging  5.7  at  Caloosahatchee  River  during  two  sampling 
periods,  which  are  from  1985  to  1989  and  from  1994  to  1995  (Doering  and  Chamberlain, 
1997).  By  contrast,  total  nitrogen  to  total  phosphorous  ratios  were  above  16,  averaging  35 
in  Caloosahatchee  River.  With  these  results,  dissolved  inorganic  N:P  ratios  (<16)  suggest 
that  nitrogen  could  limit  phytoplankton  productivity  in  agreement  with  McPherson  and 
Miller  (1990),  while  total  N:P  ration  (>16)  indicated  that  phosphorous  could  limit 
productivity  (Doering,  1996).  In  fact,  nutrient  addition  studies  conducted  by  FDER  found 
N  and  P  to  be  co-limiting  (Degrove,  1981) 

The  distributions  of  phosphorous  in  the  system  were  nearly  conservative  and  a 
function  of  river  phosphorous  concentration,  flow,  and  physical  mixing  (McPherson  and 
Millor,  1990).  A  large  amount  of  phosphorous  from  the  watershed  is  carried  by  freshwater 
discharge  into  the  tidal  reaches  of  the  Peace  River,  with  subsequent  dilution  by  the  lower- 
nutrient  seawater  entering  the  estuary  from  the  Gulf  of  Mexico.  Concentrations  of  total 
phosphorous  averaged  about  0.08  mg/L  in  Pine  Island  Sound,  0.15  mg/L  in  the  tidal 
Caloosahatchee  River,  and  0.62  mg/L  in  the  tidal  Peace  River  according  to  USGS  nutrient 
data  from  1982  to  1989  (McPherson  and  Miller,  1990). 

Most  of  the  nitrogen  in  the  rivers  and  estuary  is  organic  nitrogen  (McPherson  and 
Miller,  1990).  Organic  nitrogen  concentrations  decreased  over  the  salinity  gradient, 
indicating  river  input  as  a  source.  The  relatively  low  concentrations  of  inorganic  nitrogen 
could  limit  plant  growth  in  the  estuary  (McPherson  and  Miller,  1990).  Ratios  of  nitrogen  and 
phosphorous  constituents  at  the  Peace  River,  Myakka  River,  and  Upper  Charlotte  Harbor 
determined  by  USGS  nutrient  data  during  1975  through  1990  are  shown  in  Table  2.1  (Pribble 


IS 


etal.,  1998) 

Table  2.1  Ratios  of  nitrogen  and  phosphorous  constituents  (Pribble  et  al.,  1998) 
RATIO Peace  River     Myakka  River     Upper  Harbor 


NH3  :  TN 

0.0327 

0.0349 

0.0338 

N03 : TN 

0.4378 

0.0505 

0.2442 

ON:TN 

0.5614 

0.9096 

0.7355 

P04  :  TP 

0.9384 

0.9371 

0.9361 

OP:TP 

0.0616 

0.0683 

0.0650 

Concentrations  of  ammonia  were  highly  variable  along  the  salinity  gradient  and  were 
in  about  the  same  range  as  concentrations  in  the  rivers  (McPherson  and  Miller,  1990). 
Ammonia  concentrations  increased  in  the  deeper  water  of  Charlotte  Harbor  during  summer 
(Fraser,  1986).  Ammonia  enrichment  probably  was  related  to  density  stratification  and  to 
low  concentrations  of  dissolved  oxygen  in  bottom  waters  (McPherson  et  al,  1996). 
Concentrations  of  nitrate  and  nitrite  nitrogen  were  nonconservative  and  decreased  sharply 
along  the  salinity  gradient  (McPherson  and  Miller,  1990).  The  sharp  decline  of  the  nitrate 
and  nitrite  nitrogen  in  the  low  salinity  regions  indicates  a  substantial  removal  of  nitrogen 
from  the  water  column  due  to  biological  uptake. 
2.3.2  Dissolved  Oxygen 

Dissolved  oxygen  is  critical  for  survival  of  plants  and  animals  in  fresh  and  salt  water, 
and  a  major  constituent  of  interest  in  water  quality  study.  Dissolved  oxygen  concentrations 
in  the  near  surface  water  of  the  system  ranged  from  about  to  6  to  8  mg/L  during  daylight 
samplingin  1982 -84  (Stoker,  1986).  Dissolved  oxygen  concentrations  of  near  bottom  water 
of  the  estuary  generally  are  lower  than  near  surface  concentrations. 

Bottom  water  hypoxia  (dissolved  oxygen  <2  mg/L)  in  Charlotte  Harbor  has  been 
reported  periodically  by  the  Environmental  Quality  Laboratory  (EQL)  since  1975  (Heyl. 


19 

1996).  Dissolved  oxygen  concentrations  were  measured  monthly  from  1976  to  1984  in  the 
river  mouth  of  Peace  River  near  PuntaGorda  (CH-006)  (Fraser,  1986).  The  average  monthly 
near-surface  concentrations  declined  from  8.5  to  6.7  mg/L  from  January  to  July  and  then 
began  to  rise  (Figure  2.3).  Near-bottom  average  monthly  concentrations  at  this  area  were 
highest  in  February,  declined  slowly  through  May,  and  then  declined  more  rapidly  until  July 
(Fraser,  1986).  The  hypoxia  conditions  during  summer  are  attributed  to  strong  stratification, 
which  cause  restricted  reaeration,  and  to  SOD.  After  breakup  of  the  stratification,  the 
concentration  increased  from  October  to  December. 


F      M 


M      J 


A      S       O      N 


Figure  2.3    Average  monthly  concentration  of  dissolved  oxygen  in  upper  Charlotte 
Harbor,  site  CH-006,  1976-84  (Fraser,  1986) 

2.3.3  Phytoplankton 

Phytoplankton  is  an  important  component  of  water  quality  processes  and  a  major 
primary  producer  in  coastal  and  estuarine  waters.  The  temporal  and  spatial  variability  of 


20 
phytoplankton  productivity  and  biomass  in  Charlotte  Harbor  have  been  investigated  by 
Environmental  Quality  Laboratory,  Inc  (EQL)  (1987).  Phytoplankton  productivity  and 
biomass  (as  chlorophyll _a)  in  the  system  are  relatively  low  most  of  time.  Productivity 
ranged  from  5  to  343  (mgC/m3)/h  and  averaged  59  (mgC/m3)/h  from  1985  to  1986 
(McPherson  et  al.,  1990).  Chlorophylljx  concentrations  ranged  from  1  to  46  mg/m3  and 
averaged  8.5  mg/m3.  Both  productivity  and  biomass  were  greater  during  summer  near  the 
mouth  of  tidal  rivers  which  has  middle  range  salinity  of  6  to  12  ppt  (McPherson  et  al.,  1990). 
Phytoplankton  productivity  and  biomass  in  the  system  are  affected  by  freshwater 
inflow  that  lowers  salinity,  increases  nutrient  availability,  and  reduces  light  penetration  in  the 
water  column.  The  nutrient  rich  colored  water  is  diluted  by  seawater  at  middle  range  salinity 
of  10  to  20  ppt,  so  that  availability  of  light  increases  and  sufficient  nutrient  concentrations 
remain  available  from  runoff,  to  stimulate  productivity  and  growth  of  phytoplankton  in  these 
areas  (McPherson  et  al.,  1990).  Of  the  major  nutrients  for  phytoplankton  productivity, 
inorganic  nitrogen  is  in  lowest  supply  and  most  critical  in  limiting  phytoplankton 
productivity  and  growth  in  the  system  (Fraser  and  Wilcox,  1981). 

The  composition  of  the  phytoplankton  in  the  system  varied  with  location  and  season 
(McPherson  et  al.,  1990).  Diatoms  were  dominant  in  55  %  of  289  phytoplankton  samples 
collected  in  the  system  between  1983  and  1984,  cryptophytes  in  35  %,  cyanophytes  (blue- 
green  algae)  in  about  6  %,  dinoflagellates  in  about  4  %,  and  other  classes  in  1  %  (McPherson 
et  al.,  1996). 

2.4  Sediment 

Sediment  quality  data  from  Charlotte  Harbor  are  scarce.  Organic  carbon,  nitrogen 
and  phosphorous  data  exist  in  only  two  studies.  Hwang  (1966)  obtained  data  in  1965  from 


21 
119  stations  throughout  the  study  area.  The  FDEP  Coastal  Sediment  Contaminant  Survey 
obtained  carbon,  nitrogen  phosphorous,  and  metal  data  in  the  sediment  column  from  33 
stations  during  1985  through  1989,  in  all  areas  except  the  Gulf  of  Mexico  (Sloane,  1994). 
Mean  carbon/nitrogen  ratios  in  the  six  different  areas  of  Charlotte  Harbor  ranged  from  8.9 
at  Lower  Charlotte  Harbor  to  16.4  at  Gasparilla  Sound  (Schropp,  1998).  Despite  data 
limitation,  the  available  data  indicate  that  Charlotte  Harbor  is  relatively  free  of  sediment 
contamination  in  comparison  with  Tampa  Bay  and  Biscayne  Bay,  where  sediment 
contaminations  are  widespread  and  are  present  in  higher  concentrations  (Schropp,  1998). 

2.5  Light  Environment 

The  amount  of  photosynthetically  active  radiation  (PAR)  in  natural  water  is 
fundamentally  important  in  determining  the  growth  and  vigor  of  aquatic  plants  (McPherson 
et  al.,  1996).  Light  is  reflected,  absorbed,  and  refracted  by  dissolved  and  suspended 
substances  in  the  water  column,  and  by  water  itself.  The  controlling  factors  are  chlorophyll- 
a,  dissolved  substances,  and  non-algal  particulate  matter  (Christian  and  Sheng,  2003). 

Dissolved  and  suspended  matter  are  the  major  causes  of  light  attenuation  in  the 
system:  phytoplankton  and  chlorophyll-a  are  generally  minor  causes  of  attenuation 
(McPherson  and  Miller,  1990).  On  average,  non-chlorophyll  suspended  matter  accounted 
for  72%  of  light  attenuation,  dissolved  organic  matter  (color)  accounted  for  21%, 
phytoplankton  chlorophyll  for  4%,  and  water  for  the  remaining  3%  (McPherson  et  al.,  1996). 
Water  color  can  cause  light  attenuation  at  tidal  rivers  because  the  source  of  the  water  color 
is  terrestrial  dissolved  organic  matters.  Dissolved  organic  matter  has  little  effect  on  light 
attenuation  in  much  of  the  southern  part  of  the  estuary,  where  suspended  matter  is  the  major 
cause  of  light  attenuation  (McPherson  and  Miller,  1996).  The  source  of  suspended  matter 


22 

includes  the  bottom  of  the  estuary,  which  consisted  of  very  fine  to  fine  sand  (Hwang,  1966) 
and  organic  detrital -material  (McPherson  and  Miller,  1990),  and  the  major  rivers. 


CHAPTER  3 
HYDRODYNAMICS  AND  SEDIMENT  TRANSPORT  MODEL 

Biological  processes  in  an  estuarine  system  are  strongly  affected  by  hydrodynamics 
and  sediment  transport  processes.  Therefore,  as  the  first  step  to  study  nutrient  cycling  in  an 
estuary,  hydrodynamics  and  sediment  transport  processes  must  be  investigated  and 
understood. 

The  three-dimensional  hydrodynamics  model  CH3D  (Sheng  et  al.,  2001)  and  the 
associated  sediment  transport  model  CH3D-SED3D  (Sheng  et  al.,  2000a)  are  used  for 
numerical  simulations  in  this  study.  The  model  framework  has  been  improved  and  modified 
from  earlier  versions  in  order  to  develop  an  integrated  model  that  couples  hydrodynamics, 
sediment  and  water  quality  dynamics.  The  code  of  the  enhanced  integrated  model  is 
optimized  to  achieve  more  efficient  coupling  and  more  systematic  structure.  The  detail 
structure  of  enhanced  model  is  explained  in  Appendix  A  as  flow  chart.  The  application  of 
the  circulation  and  transport  model  to  produce  a  detailed  characterization  of  hydrodynamics 
within  system  is  the  first  step  in  the  development  of  the  integrated  model  of  the  system. 

The  CH3D  model  for  Tampa  Bay  and  Indian  River  Lagoon  did  not  include  the 
simulation  of  temperature  distribution.  In  this  study,  the  temperature  field  is  simulated  by 
solving  the  temperature  equation  with  an  air-sea  interface  boundary  condition.  Temperature 
is  an  important  factor  for  baroclinic  circulation  and  water  quality  processes,  in  which  almost 
all  reaction  parameters  in  the  nutrient  cycle  are  a  functions  of  temperature.    Therefore, 


23 


24 
improved  simulation  of  temperature  is  expected  to  produce  more  accurate  baroclinc  and 
water  quality  simulations. 

3.1  Governing  Equation 

Hydrodynamics  and  sediment  transport  processes  in  estuaries  are  complicated,  three- 
dimensional  and  time-dependent.  Mathematical  descriptions  of  these  processes  generally 
require  simplifying  assumptions. 
3.1.1  Hydrodynamic  Model 

The  governing  equations  that  describe  the  velocity  and  surface  elevation  fields  in 
shallow  water  are  derived  from  the  Navier-Stokes  equations.  In  general,  four  simplifying 
approximations  are  applied:  1)  the  flow  is  incompressible,  which  results  in  a  simplified 
continuity  equation,  2)  the  horizontal  scale  is  much  larger  than  the  vertical  scale  such  that 
the  hydrostatic  pressure  distribution  is  valid,  3)  the  Boussinesq  approximation  can  be  used 
to  simplify  the  treatment  of  the  baroclinic  terms,  4)  the  eddy-viscosity  concept,  which 
assumes  that  the  turbulent  Reynolds  stresses  are  the  product  of  mean  velocity  gradients  and 
"eddy  viscosity",  can  be  employed.  With  the  above  assumptions,  the  continuity  equation  and 
x-  and  y-  momentum  equations  have  the  following  form  (Sheng,  1983): 

du      dv     dw      _ 

—  + —  +  —  =  0  (3.1) 

dx     dy      oz 


+ 
dz 


du      duu      duv     duw  _        dC, 

dt       dx        dy        dz  dx  H \dx2      dy 

dv     dvu     dvv     dvw  dC,  ( d2v     32v|      dfA    oV 


fd2u   aV 

J 


d  (       du 


^ 


(3.2) 


UVU  UVV         UVYV  l/L,  V     V  V     r         ,      w    \      A       v  /TT\ 


dt      dx       dy       dz  '  dy  {dx2     dy 


J 


dz\       dz 


where   u(x,y,z,t),   v(x,y,z,t),   w(x,y,z,t)   are  the  velocity  components  in  the 
horizontal  x-  and  y-directions,  and  vertical  z-direction;  t  is  time;  g(x,  y,  t)  is  the  free  surface 


25 


elevation;  g  is  the  gravitational  acceleration;  and  AH  and  \  are  the  horizontal  and  vertical 

turbulent  eddy  coefficients,  respectively. 

In  Cartesian  coordinates,  the  conservation  of  salt  and  temperature  can  be  written  as 


dS     duS     dvS     dwS      d 

—  + + + =  — 

dt       dx       dy       dz       dx 
dT     duT     dvT     dwT      d 


DH  — 

ox 


d' 

+  — 
dy 


dy 


dz 


+• 


+  ■ 


■  +  ■ 


dt       dx        dy        dz 


Kt 


dT_ 
dx 


+  ■ 


<       dT 


dy[        dy 


+  ■ 


dz 


DvTz 

v    V  dz 


(3.4) 


(3.5) 


where  S  is  salinity;  T  is  temperature;   DH  and  KH    are  the  horizontal  turbulent  eddy 
diffusivity  coefficients  for  salinity  and  temperature,  respectively;  and   Dv  and  Kv  are  the 

vertical  turbulent  eddy  diffusivity  coefficients  for  salinity  and  temperature,  respectively. 

Since  the  length  scales  of  horizontal  motion  in  estuarine  systems  are  much  greater 
than  those  of  vertical  motion,  it  is  common  to  treat  vertical  turbulence  and  horizontal 
turbulence  separately.  In  shallow  estuaries,  the  effect  of  the  horizontal  eddy  viscosities  on 
circulation  are  much  smaller  than  the  effect  of  the  vertical  eddy  viscosity,  although  the 
horizontal  eddy  viscosity  is  typically  2-3  orders  of  magnitude  larger  than  the  vertical  eddy 
viscosity. 

Vertical  turbulent  mixing  is  an  important  process,  which  can  significantly  affect 
circulation  and  transport  in  an  estuary.  Since  turbulence  is  a  property  of  the  flow  instead  of 
the  fluid,  it  is  essential  to  use  a  robust  turbulence  model  to  parameterize  the  vertical  turbulent 
mixing.  In  this  study,  the  vertical  eddy  coefficients  (  Av ,  Dv  and  Kv )  are  computed  from 

a  simplified  second-order  closure  model  developed  by  Sheng  and  Chiu  (1986),  and  Sheng 
and  Villaret  (1989). 


26 

Various  form  forms  of  the  equation  of  state  can  be  used.  The  present  model  uses  the 
equation  given  by  (Eckart,  1958) 

P 

P  = 


(a  +  0.698P) 

P  =  5890  + 387/ -0.357r2  +  3S  (3.6) 

a  =  1779.5  +  1 1.257  -0.04757/2-  (3.8  +  0.017)5 

where  T  is  in  °C  ,  S  is  in  ppt  and  p  is  in  g  I  cm  . 

The  complete  details  of  model  equations  in  the  curvilinear  boundary-fitted  and  sigma 
coordinates  have  been  derived  (Sheng,  1989),  and  are  presented  in  Appendix  B. 
3.1.2  Sediment  Transport  Model 

The  suspended  sediment  model  includes  the  advection-diffusion  processes,  which  are 
computed  by  the  hydrodynamics  model,  as  well  as  such  processes  as  erosion,  deposition, 
flocculation,  settling,  consolidation,  and  entrainment  (Sheng,  1986;  Metha,  1986). 

The  governing  equation  that  represents  the  transport  of  suspended  sediments  is  given 
by: 


dc     due     dvc     d(w  +  w)c      d 

—  + +  ^  + 


dt      dx       dy  dz  dx  \    H  dx )     dy\       dy 


B* 


+  ■ 


'B^ 


K      dz 


c)z 


(3.7) 


where  c  is  the  suspended  sediment  concentration,  w5  is  the  settling  velocity  of  suspended 
sediment  particles  (negative  downward),  BH  is  the  horizontal  turbulent  eddy  diffusivity,  and  Bv 

is  the  vertical  turbulent  eddy  diffusivity. 

Four  simplifying  approximations  are  implied  in  the  above  equation:  1)  the  concept 
of  eddy  diffusivity  is  valid  for  the  turbulent  mixing  of  suspended  sediments,  2)  the  suspended 
sediment  dynamics  are  represented  by  the  concentrations  of  two  particle  size  groups  (Sheng 
et  al.,  2002a),  3)  the  suspended  sediment  concentrations  are  sufficiently  small  that  all 


27 
particles  follow  turbulent  eddy  motions,  4)  the  SSC  is  sufficiently  low  that  non-Newtonian 
behavior  can  be  neglected. 

In  this  study,  the  determination  of  settling,  flocculation,  deposition,  erosion, 
flocculation,  and  consolidation  processes  are  based  on  the  previous  work  of  Sheng  and  Lick 
(1979),  Sheng(1986),  Metha  (1989),  Sheng  et  al.  (1990),  Chen  and  Sheng  (1994)  and  Sun 
and  Sheng  (2001). 

3.2  Boundary  and  Initial  Conditions 
Obtaining  solutions  for  Equations  (3.1)  Through  (3.7)  requires  the  specification  of 
appropriate  boundary  and  initial  conditions. 
3.2.1  Boundary  Conditions 

The  boundary  conditions  at  the  free  surface  (a=0)  in  a  non-dimensional,  vertically 
stretched,  boundary-fitted  coordinate  system  are: 

du       H     „ 


For  Hydrodynamics:  \ 


dcr     Ey 


*"  3.7     E,    " 

For  Salinity:  =  0  (3.8) 

dcr 


dT      H  Pr, 


v 


For  Temperature:  — —  = qT 


dcr        E 


v 


Dv  fci     ^ 

For  Sediment:  w.c.  +  —^—^  =  0 

"  '      H  dcr 
where  Ev  is  a  vertical  Ekman  number  and  Prv  is  a  vertical  Prandtl  number;  The  Cartesian 

wind  stress,  t",  is  calculated  using 

Tw  =pC,v  Ju2  +v2  (3.9) 

y         fa     as    w  v     w  w  v         ' 


28 
where  pa  is  the  air  density  (0.0012  g/cm3),  uw  and  vw  are  the  components  of  wind  speed 
measured  at  some  height  above  the  sea  level.  Cds,  the  drag  coefficient,  is  given  as  a  function 
of  the  wind  speed  measured  at  10  meters  above  the  water  surface  by  (Garrat,  1977). 

Cds  =  0.001(0.75  +  0.067  Ws)  (3.10) 

Boundary  conditions,  if  specified  in  a  Cartesian  coordinate  system,  such  as  wind 
stress,  must  be  transformed  before  being  used  in  the  boundary-fitted  equations.  For  example, 
the  surface  stress  in  the  transformed  system  is  given  by 


3£        3£ 
dx  dy 

drj         drj 


r<-ar-+*r< 


r,„=-^,+— r,v  (3-u) 

ox        oy 

The  second  surface  boundary  condition  is  the  kinematic  free  surface  boundary 
condition  which  states 

BC       d£      dC 

w  =  -?-  +  u-^  +  v^f-  (3.12) 

dt         ox        oy 
The  boundary  conditions  at  the  bottom  in  a  non-dimensional,  vertically  stretched  (o 

=  -1),  boundary-fitted  coordinate  system  are 

For  Hydrodynamics: 

\  -£-  =  —T^  =-^HrZrCd  [guu2b  +  2gl2ubvb  +  g22v;~\ub 


da     Ev    "'      A, 

A  ^L  =  JLt-=£j-HZC 

da     Ev  \r 


guu;  +  2gnubvb  +  g22v2b~\vb 


For  Salinity:  ^  =  °  (313) 

dcr 

dT      n 
For  Temperature:  — —  =  U 

da 

Dv  dc 

For  Sediment:  w.ci  +  — -— -1-  =  A  _  Et 

si  '       H  da         '        ' 


29 
where  ub  and  vh  are  the  contravarient  velocity  components  at  the  first  grid  point  above  the 
bottom.  Ej  is  the  erosion  rate  and  D,  is  the  deposition  rate  for  sediment  group  i.  Cd  is  drag 
coefficient  which  is  a  function  of  the  size  of  bottom  roughness  elements,  z0,  and  the  height 
at  which  uh  is  measured,  so  long  as  z,  is  within  the  constant  flux  layer  above  the  bottom.  The 
drag  coefficient  is  given  by  (Sheng,  1983) 


cd  = 


K 


(3.14) 


ln(z,/z0) 
where  k=0.4  is  the  von  Karman  constant,  z0=k/30  and  ks  is  the  bottom  roughness. 

Along  the  shoreline  where  river  inflow  may  occur,  the  conditions  are  generally 

u  =  u(x,y,(J,t) 

v  =  v(x,y,cr,t) 

w  =  0 

(3.15) 
S  =  S(x,y,cr,t) 

T  =  T(x,y,cr,t) 

c,  =ci(x,y,a,t) 

Contravarient  velocity  components  provide  lateral  boundary  conditions  similar  to 

those  in  Cartesian  systems  (x,y).  Along  solid  boundaries,  the  normal  velocity  component 

must  be  zero  to  satisfy  the  no-slip  condition.   In  addition,  normal  derivatives  of  salinity, 

temperature  and  suspended  sediment  concentration  are  assumed  to  be  zero.  When  flow  is 

specified  at  a  boundary,  the  normal  velocity  component  is  prescribed.    Tidal  boundary 

conditions  are  specified  using  water  level,  (,  directly  When  tidal  boundary  conditions  are 

given  in  terms  of  C  the  normal  velocity  component  is  assumed  to  be  of  zero  slope,  while 

tangential  velocity  component  may  be  either  zero,  of  zero  slope,  or  computed  from  the 

momentum  equations.  During  an  ebb  tide,  the  concentrations  of  salinity,  temperature,  and 

suspended  sediment  flowing  out  are  calculated  using  a  1-D  advection  equation  while  during 


30 
a  flood  tide  the  offshore  concentrations  are  generally  prescribed  as  either  fixed  or  time 
varying. 
3.2.2  Initial  Conditions 

To  initiate  a  simulation,  the  initial  spatial  distribution  of  (,,  u,  v,  w,  S,  T,  and  c,  must 
be  specified.  When  these  values  are  unknown,  "zero"  initial  fields  can  be  used.  When  these 
values  are  known  at  a  limited  number  of  locations,  an  initial  field  can  be  generated  by  spatial 
interpolation.  In  the  principle,  the  interpolated  field  should  satisfy  the  conservation  equation 
for  a  particular  variable.  For  practical  simulations,  a  "spin-up"  period  is  required  to  damp 
out  transients  caused  by  the  initial  condition,  which  does  not  satisfied  conservation.  The 
length  of  a  spin  up  period  is  variable  and  depends  on  such  factors  as  basin  size,  flushing 
time,  and  current  velocity. 

3.3  Heat  Flux  at  the  Air-Sea  Interface 

The  ocean  receives  energy  through  the  air-sea  interface  by  exchange  of  momentum 
and  heat.  The  temperature  of  ocean  waters  varies  from  place  to  place  and  from  time  to  time. 
Such  variations  are  indications  of  heat  transfer  by  currents,  absorption  of  solar  energy,  and 
loss  by  evaporation.  The  size  and  character  of  the  temperature  variations  depends  on  the  net 
rate  of  heat  flow  into  or  out  of  water  body  (Pickard  and  Emery,  1990).  The  transfer  of  heat 
across  the  air-sea  interface  determines  the  distribution  of  temperature  in  the  ocean  as  a 
surface  boundary  condition  for  the  temperature  equation,  as  follow: 

dT 
PoKv  —  =  qT  (3-16) 

az 

where  qT  is  the  net  heat  flux  across  the  surface  water. 

To  estimate  the  net  heat  flux  at  the  air-sea  interface,  it  is  necessary  to  consider  the 
following  processes:  net  heat  flux  across  the  surface,  qT\  heating  by  incoming  solar  radiation 


31 

(insolation),  qs  ;  warming  by  conductive  heat  exchange  across  the  surface  (sensible  heat 
flux),  qh  ;  cooling  by  net  outgoing  long  wave  radiation,  qh ;  heat  loss  as  water  evaporated 
(latent  heat  flux),  q, ;  heat  flux  from  precipitation.  Other  source  of  heat  flow,  such  as  that 
from  the  earth's  interior,  change  of  kinetic  energy  of  waves  into  heat  at  the  surf;  heat  from 
chemical  or  nuclear  reactions,  are  small  and  can  be  neglected.  The  net  heat  exchange  at  the 
surface  can  be  divided  into  four  terms: 

qT  =  <2s-<ib-<ik-<ii  (3-17) 

Direct  measurement  of  these  fluxes  is  the  best  way  to  provide  the  net  heat  flux  at  the 
air-sea  interface.  However,  directly  measured  air-sea  fluxes  are  only  available  at  very  few 
stations  to  allow  calculation  of  air-sea  interface  over  a  large  area.  Instead,  directly  measured 
air-sea  fluxes  are  used  for  developing,  calibrating,  and  verifying  the  parametric  formulae  for 
estimating  the  fluxes  from  the  primary  variables  including  wind  speed,  air  temperature,  water 
temperature,  and  cloud  cover  (Taylor  et  al.,  2000).  These  required  parametric  formulae  are 
then  used  to  compute  heat  fluxes  over  a  large  area. 
3.3.1  Short- Wave  Solar  Radiation 

The  main  source  of  heat  flux  through  the  air-sea  interface  is  short-wave  solar 
radiation,  gv  (incoming  solar  radiation),  received  either  directly  or  by  reflection  and  scattering 
from  clouds  and  the  atmosphere.  The  incoming  solar  radiation  is  based  on  the  empirical 
formula  of  Reed  (1977): 

qs  =  S  sin  y  ( A  +  fisin  y)(\  -  0.62n  -  0.0019r)(l  -  a)  (3.18) 

where  S  is  solar  constant  =1353  w/m2 
Y  is  solar  elevation  angle 
n  is  cloud  cover  =  0-1 


32 

a  is  the  albedo  =  0.06 

A  and  B  are  empirically  determined  coefficients  for  each  category  of  reported  total 
cloud  amount  and  the  solar  elevation 

The  solar  elevation  angle  was  computed  to  the  nearest  0. 1  °  from  date  and  time  for  the 

latitude  and  longitude  of  the  study  area,  using  the  following  equations  (Miller  and 

McPherson,  1995): 

360 
<p  =  (d-l) 

365.242 

J  =  12  +  0.1236sin^-0.0043cos^  +  0.1538sin2^  +  0.0608cos2^ 
<t  =  279.9348  +  p  + 1.9148  sin  p-0.0795cos^  + 0.00199  sin  2^  + O.OO16cos20 

Y  =  \5(t-S)-A  (3.19) 

k  =  arcsin  (0.39785077  x  sin  a) 
sin  /3  -  sin  /sin  k  +  cos  y  cos  /f  cos  Y 

where  (p  is  the  angular  fraction  of  the  year,  in  degree;  d  is  Julian  date;  6  is  true  solar  noon, 
in  hours;  T  is  the  solar  hour  angle,  in  degrees;  T  is  the  Greenwich  Mean  Time,  in  hours;  X 
is  the  longitude,  in  degrees;  o  is  an  estimate  of  the  true  longitude  of  the  sun,  in  degrees;  K  is 
the  solar  declination,  in  degrees;  y  is  the  latitude,  in  degrees;  and  P  is  the  solar  elevation 
angle,  in  degrees. 
3.3.2  Long-Wave  Solar  Radiation 

The  back  radiation  term,  qh,  is  the  net  amount  of  energy  lost  by  the  sea  as  long-wave 
radiation.  The  amount  of  long  wave  flux  is  dependent  on  surface  water  temperature, 
atmospheric  temperature,  humidity  and  cloud  cover  (Clark  et  al,  1974) 

qb  =  eoT*  (0.39  -  0.05e05 )  (l  -  An2 )  +  4eaT*  (Ts  -  Ta )  (3.20) 

where  e  is  the  emittance  of  the  sea  surface  =  0.98 

o  is  Stefan-Boltzman  constant  =  5.673  x  10"8  W/m2 


33 

e  is  the  water  vapor  pressure 
Ta  and  Ts  are  the  air  and  sea  temperatures  in  K. 
A  is  a  cloud  cover  coefficient  which  varies  with  latitude 

The  Antoine  constants  (http://www.owlnet.rice.edu/~ceng301/31.html)  give  the  vapor 
pressure  as  a  function  of  temperature  so  that  the  approximate  value  for  water  vapor  pressure 


is: 


(  3985.44    A 

16.5362- 


e  =  exp 

Ts  -38.997 

where  Ts  is  the  surface  water  temperature  (°K). 


(3.21) 


A  theoretical  calculation  of  the  mean  values  of  the  coefficient  X  for  different  latitudes 
has  been  made  by  M.  E.  Berliand  (1952).  In  this  calculation,  the  mean  frequency  of  clouds 
of  different  layers  at  each  latitude  is  taken  into  account.  The  values  obtained  for  the 
coefficient  X  are  given  in  Table  3.1  (Budyko,  1974).  The  reduction  of  values  of  this 
coefficient  in  low  latitudes  is  explained  mainly  by  a  higher  mean  altitude  of  clouds  in  these 
regions. 
Table  3.1  Mean  latitudinal  values  of  the  coefficient  X 


* 

75° 

70° 

65° 

60° 

55° 

50° 

45° 

40° 

X 

0.82 

0.80 

0.78 

0.76 

0.74 

0.72 

0.70 

0.68 

d> 

35° 

30° 

25° 

20° 

15° 

10° 

5° 

0° 

X 

0.65 

0.63 

0.61 

0.59 

0.57 

0.55 

0.52 

0.50 

3.3.3  Sensible  and  Latent  Heat  Fluxes 

The  most  significant  part  in  terms  of  heat  transfer  from  the  sea  to  the  atmosphere  is 
latent  heat  flux.  The  rate  of  heat  loss  is  equal  to  the  rate  of  vaporization  times  the  latent  heat 
of  evaporation.  Sensible  heat  flux  is  due  to  temperature  gradient  in  the  air  above  the  sea. 
The  rate  of  loss  or  gain  of  heat  is  proportional  to  the  temperature  gradient,  heat  conductivity 


34 
and  the  specific  heat  of  air  at  constant  pressure.  Sensible  and  latent  heat  fluxes  are  estimated 
by  using  the  bulk  aerodynamic  equations  (Mellor,  1996): 

qk=pcpCHUw(Ts-Ta)  (3.22) 

qi=pCELUl0(hs-ha)  (3.23) 

where  p  is  the  density  of  air 

CH  and  CE  are  transfer  coefficients  for  sensible  and  latent  heat  respectively 

UI0  is  the  wind  speed  at  10  m  height  above  surface 

Ts  and  Ta  are  the  sea  surface  and  air  temperatures 

cp  is  the  specific  heat  of  air,  1.0048  x  103  J/kg°K 

L  is  the  latent  heat  of  evaporation,  2.4  x  106  J/kg 

hs  and  ha  are  specific  humidities  at  the  surface  and  at  a  10  m  height  above  the  surface 

The    air   density   is   determined   from   the   ideal    gas   equation   in   the   form, 

p  =  pJ(RTa ) ,  where  Pa  is  air  pressure  and  R  is  universal  gas  constant,  287.04  J/kg  °K. 

The  traditional  estimates  of  CH  and  CE  over  the  ocean  tend  to  support  a  fairly  constant  value 
over  a  wide  range  of  wind  speed.  Smith  (1989)  recommended  a  constant  "consensus"  value 
CE  =  (1.2  ±  0.1)xl0"3  for  winds  between  4  and  14  m/s.  DeCosmo  et  al.  (1996)  also  suggested 
a  near  constant  value  with  CE  =  (1.12  ±  0.24)xl0 3  for  winds  up  to  18  m/s.  For  the  Stanton 
number,  CH,  Friehe  and  Schmitt  (1976)  obtained  slightly  different  values  for  unstable  and 
stable  conditions,  0.97  xlO"3  and  0.86  xlO'3  respectively.  Smith  (1989)  suggested  CH  =  1.0 
x  10 3.    CE-  1.2  xlO"3 and  CH  =  1.0  x  10"3  were  used  for  this  study. 

The  formulae  used  for  air  specific  heat,  cp  ,  and  latent  heat  for  evaporation,  L,  have 
been  taken  from  Stull  (1988), 

cp  =1004.67  +  0.84^  (3.24) 


35 

L  =  106  [2.501  -0.00237(7;  -273.16)]  (3.25) 

An  empirical  relation  (based  on  the  Clausius-Clapeyron  theory)  provides  the 

•    i  ir.(0.7859+O.O3477r)/(l+O.0O4l2r)        ,  »      „-„_-„! 

saturation,  water  vapor  partial  pressure,  es  =  10v  mb.    A  general 

relation  between  specific  humidity  and  partial  pressure  is  q  -  (Q.622e/P)/(l  -  0.31Se/P) 

,  where  P  is  the  atmospheric  pressure  (Mellor,  1996).  Specific  humidities  at  surface  and  at 
a  10  m  height  above  the  surface,  hs  and  ha,  can  be  calculated  using  these  empirical  formulae 
with  Ts  and  Ta  respectively. 


CHAPTER  4 
WATER  QUALITY  MODEL 

In  this  chapter,  a  water  quality  model  for  simulating  water  quality  processes  occurring 
in  the  Charlotte  Harbor  estuarine  system  in  both  the  water  and  sediment  columns  is 
presented.  This  is  an  extension  of  an  earlier  model  developed  to  simulate  water  quality  for 
Lake  Okeechobee  (Sheng  et  al.,  1993;  Chen  and  Sheng,  1994),  Tampa  Bay  (Yassuda  and 
Sheng,  1996)  and  the  Indian  River  Lagoon  (Sheng  et  al.,  2001  and  2002).  These  models 
include  the  effect  of  sediment  transport  on  nutrient  dynamics  through  the  explicit  use  of  a 
sediment  transport  model  and  the  incorporation  of  a  nutrient  resuspension  flux. 

The  water  quality  model  incorporates  on  the  interactions  between  oxygen  balance, 
nutrient  dynamics,  light  attenuation,  temperature,  salinity,  phytoplankton  and  zooplankton 
dynamics.  To  develop  the  water  quality  model,  the  mass  conservation  principle  can  be 
applied  to  each  water  quality  parameter,  as  it  relates  to  phytoplankton  and  zooplankton 
dynamics,  nitrogen  and  phosphorous  cycles,  and  oxygen  balance. 

The  nitrogen  and  phosphorous  cycles  in  an  estuarine  system,  are  modeled  through  a 
series  of  first  order  kinetics.  Nutrient  concentrations  in  the  estuarine  system  are  constantly 
changing  in  time  and  space  due  to  loading  from  rivers,  exchanges  with  the  ocean,  seasonal 
climatic  changes,  biogeochemical  transformations,  hydrodynamics,  and  sediment  dynamics. 
Nitrogen  species  include  ammonia  nitrogen  (NH3),  soluble  ammonium  nitrogen  (NH4), 
nitrate  and  nitrite  nitrogen  (N03),  particulate  ammonium  nitrogen  (PIN),  soluble  organic 


36 


37 

nitrogen  (SON),  particulate  organic  nitrogen  (PON),  phytoplankton  nitrogen  (PhyN),  and 
zooplankton  nitrogen  (ZooN).  Similar  to  the  nitrogen  species,  phosphorous  species  include 
soluble  reactive  phosphorous  (SRP),  particulate  inorganic  phosphorous  (PIP),  soluble  organic 
phosphorous  (SOP),  particulate  organic  phosphorous  (POP),  phytoplankton  phosphorous 
(PhyP),  and  zooplankton  phosphorous  (ZooP). 

Phytoplankton  kinetics  are  the  central  part  of  this  water  quality  model,  since  the 
primary  water  quality  issue  in  the  estuarine  system  is  eutrophication  (Boler  et  al.,  1991). 
Phytoplankton  population  is  a  complex  variable  to  measure  in  the  field.  However,  the  lack 
of  data  on  each  specific  species  prevented  a  more  detail  characterization,  the  entire 
phytoplankton  community  is  represented  in  this  study  by  a  single  state  variable  and 
quantified  as  carbonaceous  biomass.  Chlorophyll _a  concentrations,  for  comparison  with 
observations,  are  obtained  through  division  of  computed  carbonaceous  biomass  by  the 
carbon-to-chlorophyll_a  ratio. 

The  oxygen  balance  couples  dissolved  oxygen  to  other  state  variables.  Reaeration 
through  the  air-sea  interface,  and  phytoplankton  production  during  photosynthesis  are  the 
main  source  for  oxygen.  Oxidation,  nitrification,  respiration,  mortality  and  SOD  reduce 
oxygen  in  the  system.  Oxidation  of  organic  matter  and  carbonaceous  material,  respiration  by 
zooplankton  and  phytoplankton,  and  oxygen  consumption  during  the  nitrification  process  are 
collectively  grouped  into  the  CBOD  (Carbonaceous-Biogeochamical  Oxygen  Demand) 
variable,  which  is  a  sink  for  dissolved  oxygen  (Ambrose  et  al.,  1994). 

The  methods  of  coupling  with  hydrodynamic  and  sediment  transport  models,  the 
simulated  parameters,  the  assumptions,  the  chemical/biological  processes  of  the  CH3D  water 
quality  model  (CH3D-WQ3D)  was  compared  with  those  for  existing  water  quality  models, 


38 
specially  Water  Quality  Analysis  ans  Simulation  Program  (WASP),  the  integrated 
Compartment  water  quality  model  developed  by  the  US  Army  Corps  (CE-QUAL-ICM)  in 
Appendix  C. 

Temperature,  salinity,  and  light  are  important  parameters  that  effect  the  rate  of 
biogeochemical  reactions.  Most  of  the  transformation  processes  in  the  nutrient  cycle  are 
affected  by  temperature.  Salinity  also  influences  DO  saturation  concentration  and  is  used  in 
the  determination  of  kinetics  constants  that  differ  in  saline  and  fresh  water.  Light  intensity 
affects  the  photosynthesis  process  and  thus  algae  growth  rate. 

The  water  quality  model  is  enhanced  to  include  a  reaeration  model  and  a  sediment 
oxygen  demand  (SOD)  model  and  newly  coupled  with  temperature  model  and  physics-based 
light  attenuation  model.  Applying  enhanced  water  quality  model  and  a  more  accurate  light 
model  and  temperature  model  could  improve  the  vertical  distribution  and  daily  fluctuation 
of  both  phytoplankton  and  dissolved  oxygen.  Furthermore,  the  bottom  water  hypoxia  at 
upper  Charlotte  Harbor  which  is  caused  by  SOD  and  vertical  stratification  can  be  reproduced 
and  analyzed  in  this  study. 

4.1  Mathematical  Formulae 

The  water  quality  equations  are  derived  from  a  Eulerian  approach,  using  a  control 
volume  formation.  In  this  method,  the  time  rate  of  the  concentration  of  any  substance  within 
this  control  volume  is  the  net  result  of  (i)  concentration  fluxes  through  the  sides  of  the 
control  volume,  and  (ii)  production  and  sink  terms  inside  the  control  volume.  The 
conservation  equation  for  each  of  the  water  quality  parameters  is  given  by: 


M  +  V.(fl«)  =  V-[DV(fl«)]       +     Q, 
(i)  («)  (Hi)  (iv) 


(4.1) 


39 

where  (I)  is  the  evolution  term  (rate  of  change  of  concentration  in  the  control  volume),  (ii) 
is  the  advection  term  (fluxes  into/out  of  the  control  volume  due  to  the  advection  of  the  flow 
field),  (iii)  is  the  diffusion  term  (fluxes  into/out  of  the  control  volume  due  to  turbulent 
diffusion  of  the  flow  field),  and  (iv)  is  the  sink  and  source  term,  representing  the  kinetics  and 
transformations  due  to  sorption/desorption,  oxidation,  excretion,  decay,  growth,  and  bio- 
degradation.  In  the  finite  difference  solution  of  the  water  quality  model,  the  advection  and 
horizontal  diffusion  terms  are  treated  explicitly,  whereas  the  vertical  diffusion  and 
biogeochemical  transformations  are  treated  implicitly.  The  detail  descriptions  of  numerical 
solution  technique  are  described  in  Appendix  D.  The  water  quality  equations  in  the 
curvilinear  non-orthogonal  boundary  fitted  system  (£,  X],  o)  are  given  by: 


dt 


HSCV  da 


3  <    a*  ^ 


V 


'V     -s 

da 


-R, 


d(H(D)(pt 
da 


/?„ 


**\*Z 


—(yfg~0Hu<pi)  +  —  (Jg~0Hv(pi) 


+ 


^CHyJSo 


^ 


goH8 


+  - 


^CH  V  #0 


drj 


8oHS 


drj 
21  d<P, 


+  Jg0Hgl 


+  JgoHg' 


drj 


(4.2) 


J] 


+  Q, 


where  (pi   represents  any  water  quality  parameter,  -^g^    x%  me  Jacobian  of  horizontal 


transformation,  (g    ,g    ,g     )  are  the  metric  coefficients  of  coordinate  transformation, 


and  Q.  represent  biogeochemical  processes.  Equation  (4.2)  is  in  dimensionless  form  and  the 


dimensionless  constants  are  defined  in  Appendix  B. 

In  the  following  sections,  the  biogeochemical  processes  controlling  the  sink/source 
term  of  Equation  (4.2)  will  be  discussed  in  detail  for  nutrient  dynamics,  zooplankton  and 


40 
phytoplankton  dynamics,  and  oxygen  balance  in  system. 

4.2  Phytoplankton  Dynamics 

The  overall  water  quality  in  the  system  is  markedly  influenced  by  the  dynamics  of 
zooplankton  and  phytoplankton  communities  (Boleretal.,  1991).  Phytoplankton  dynamics 
and  nutrient  dynamics  are  closely  linked,  since  nutrient  uptake  during  phytoplankton  growth 
is  the  main  process  to  remove  dissolved  nutrients  from  the  water,  and  phytoplankton  and 
zooplankton  respiration  and  mortality  are  major  components  of  nutrient  recycling.  The 
diurnal  variation  of  dissolved  oxygen  is  related  with  photosynthetic  oxygen  production 
during  the  day  and  oxygen  consumption  due  to  phytoplankton  respiration  during  the  night. 
4.2.1  Modeling  Approach 

Phytoplankton  kinetics  are  represented  by  growth,  respiration,  non-predator-mortality, 
grazing  by  zooplankton,  and  a  settling  term.  The  phytoplankton  sources  and  sinks  in  the 
conservation  equation  can  be  written  as: 

dPhvC     (  d  s\ 

^1=    Ma-K^-Kas+-WSaigae    -PhyC-^ZooC  (4.3) 

at         \  az  J 

Where  PhyC  is  phytoplankton  biomass,  expressed  as  carbon  (gCm3);  /Ja  is  phytoplankton 
growth  rate  (1/d);  Kas  is  respiration  rate  (1/d);  K^  is  non-predator  mortality  (1/d);  WSalgae  is 
the  phytoplankton  settling  velocity(m/d);  ZooC  is  zooplankton  biomass  (gCm3);  and /uz  is 
zooplankton  growth  rate  (1/d). 

Phytoplankton  growth  is  determined  by  the  intensity  of  light,  by  the  availability  of 
nutrients,  and  by  the  ambient  temperature.  Light  limitation  is  formulated  according  to  the 
photo  inhibition  relationship  (Steel,  1965).  The  quantity  of  the  growth  limitation  factor  for 
nutrients  is  related  with  a  half-saturation  constant.  The  half-saturation  constant  refers  to  the 
concentration  of  the  nutrient  at  which  the  growth  rate  is  one  half  its  maximum  value.  This 


41 
results  in  a  hyperbolic  growth  curve.  An  exponential  increasing  function  is  applied  for 
temperature  limitation.. 

The  minimum  formation  approach  has  been  used  to  combine  the  limiting  factor  of 
light  and  each  limiting  nutrient.  The  minimum  formation  is  based  on  "Liebig's  law  of  the 
minimum"  which  states  that  the  factor  in  shortest  supply  will  control  the  growth  of  algae 
(Bowie  et  al.,  1980) 

Va=(Ma)aaK-f(T)'M)'f(N,P) 

(4.4) 


=  (/0™C2°-min 


— exp 
L 


Is  J 


NH4  +  N03  SRP 


Hn+NH4  +  N03    Hp+SRP 

where  OuJmax  is  the  phytoplankton  maximum  growth  rate  (1/day);  d  is  temperature 
adjustment  coefficient;  Tis  temperature  (°C);  /is  the  light  intensity,  calculated  by  the  light 
attenuation  model;  Is  is  the  optimum  light  intensity  for  algae  growth;  Hn  is  half  saturation 
concentration  for  nitrogen  uptake  (gN  — 3);  Hp  is  half  saturation  concentration  for 
phosphorous  uptake  (gP  — 3);  NH4  is  ammonium  concentration  (gN  —  3);  N03  is  nitrate 
concentration  (gN  ~~ 3);  and  SRP  is  soluble  reactive  phosphorous  concentration  (gP  — 3). 

Respiration  and  mortality  are  considered  to  be  an  exponentially  increasing  functions 
of  temperature: 

T  T  (4.5) 

where  (KaJTr  and  (KM)Tr  are  respiration  and  mortality  rate  at  Tr  (1/day);  and  Tr  is  reference 
temperature  of  respiration  and  mortality. 

For  phytoplankton,  literature  values  of  algae  settling  velocity,  which  account  for  the 
limited  vertical  motion  of  these  organisms  will  be  used. 

Zooplankton  are  included  in  water  quality  models  primarily  because  of  their  effects 


42 
on  algae  and  nutrients.  Phytoplankton  and  zooplankton  dynamics  are  closely  tied  through 
predator-prey  interaction.  Phytoplankton  dynamics  are  of  major  concern  in  this  study  while 
no  attempt  is  made  to  investigate  zooplankton  dynamics  due  to  lack  of  zooplankton  data. 
Zooplankton  is  only  considered  as  the  predators  of  phytoplankton,  utilizing  their  available 
biomass  as  food  supply.  Zooplankton  kinetics,  influenced  by  growth,  respiration  and 
mortality,  are  represented  in  a  source  and  sink  term  as  (Bowie,  1985): 

l^-fa-K.-K.yZooC  (4.6) 

at 

where  //,  is  zooplankton  growth  rate  (1/day);  Kas  is  respiration  rate  (1/day);  and  K^  is 

mortality  (1/d) 

Zooplankton  growth  is  represented  by  a  temperature-dependent  maximum  growth 

rate,  which  is  limited  by  phytoplankton  availability: 

K     '~     '        Hphr+PhyC-Trsph:1 

where  (/uz)m.M  is  the  zooplankton  maximum  growth  rate  (1/day);  6\s  temperature  adjustment 
coefficient;  Hphy  is  half  saturation  concentration  for  phytoplankton  uptake  (gC  — 3);  and  Trsphy 
is  threshold  phytoplankton  concentration  for  zooplankton  uptake  (|ig/l). 
4.2.2  Relationship  between  Phytoplankton  and  Nutrients 

Phytoplankton  biomass  is  quantified  in  units  of  carbon.  In  order  to  express  the  effects 
of  phytoplankton  on  nitrogen  and  phosphorous,  the  ratio  of  nitrogen-to-carbon  and 
phosphorous-to-carbon  in  phytoplankton  biomass  must  be  specified.  Global  mean  values  of 
these  ratios  are  well  known  (Redfield  et  al.,  1966).  The  amounts  of  nitrogen  and 
phosphorous  incorporated  in  algae  biomass  is  quantified  through  a  stoichiometric  ratio. 
Thus,  total  nitrogen  and  total  phosphorous  in  the  model  are  expressed  as: 


43 

TotN  =  NH.  +  NH,  +  NO,  +  SON  +  PON  +  PIN  +  Anc  ■  PhyC  +  Anc  ■  ZooC 

4  J  (4.8) 

TotP  =  SRP  +  SOP  +  POP  +  PIP  +  Ape  ■  PhyC  +  Ape  ■  ZooC 

where  TotN  is  total  nitrogen  (gN  ~3);  NH4  is  dissolved  ammonium  nitrogen  (gN  ~3);  NH3 
is  ammonia  nitrogen  (gN  ~3);  N03  is  nitrate  and  nitrite  nitrogen  (gN  ~3);  SON  is  soluble 
organic  nitrogen  (gNm~3);  PON  is  particulate  organic  nitrogen  (gN  _3);  PEN  is  particulate 
inorganic  nitrogen  (gN  ~3);  Anc  is  Algae  nitrogen-to-carbon  ratio  (gN/gC)  ;  TotP  is  total 
phosphorous  (gP  ~3);  SRP  is  soluble  reactive  phosphorous  (dissolved  phosphate)  (gPrn 3); 
SOP  is  soluble  organic  phosphorous  (gP-3);  POP  is  particulate  organic  phosphorous  (gP-3); 
PIP  is  particulate  inorganic  phosphorous  (gP  _3);  and  Ape  is  algae  phosphorous-to  carbon 
ratio  (gP/gC). 

The  connection  between  the  carbon,  nitrogen  and  phosphorous  cycle  is  shown  in 
Figure  4.1.  Phytoplankton  uptakes  dissolved  ammonium  nitrogen,  nitrate  and  nitrite 
nitrogen,  and  soluble  reactive  phosphorous  during  production  and  releases  dissolved 
ammonium  nitrogen,  soluble  reactive  phosphorous  and  organic  nitrogen,  organic 
phosphorous  during  respiration  and  mortality  processes.  Zooplankton  has  similar  kinetic 
processes  as  phytoplankton.  The  measured  phytoplankton  as  algae  mass  per  volume,  was 
converted  to  phytoplankton  carbon  with  a  algae  to  carbon  ratio.  The  amounts  of  nitrogen  and 
phosphorous  from  phytoplankton  can  be  converted  with  a  nitrogen  to  carbon  ratio  and  a 
phosphorous  carbon  ratio,  respectively. 

4.3  Nutrient  Dynamics 

Nutrients  are  essential  elements  for  life  processes  of  aquatic  organisms.  Nutrients 
of  concern  include  carbon,  nitrogen,  phosphorous,  silica  and  sulfur.  Among  these  nutrients, 
the  first  three  elements  are  utilized  most  heavily  by  zooplankton  and  phytoplankton.  Since 
carbon  is  usually  available  in  excess,  nitrogen  and  phosphorous  are  the  major  nutrients 


44 
regulating  the  ecological  balance  in  an  estuarine  system.  Nutrients  are  important  in  water 
quality  modeling  for  several  reasons.  For  example,  nutrient  dynamics  are  critical 
components  of  eutrophication  models  since  nutrient  availability  is  usually  the  main  factor 
controlling  algae  bloom.  Algae  growth  is  typically  limited  by  either  phosphorous  or  nitrogen 
(Bowie  et  al.,  1980).  Details  on  the  nutrient  dynamics  including  all  the  equations  used  by 
the  water  quality  model  to  calculate  the  nitrogen  and  phosphorous,  can  be  found  in 
Appendix-E. 


Excretion 


NH3 


Volatilization 


I        Sorption/ 

Desorption  w   i  r   1 1 


Mortality 


Ammonification 

*■ 


SON 


Sorption/ 


Desorption 


NITROGEN 
A 


Nitrification 


NH4 


N03 


PhyN=ar  ,'PhyC 


CARBON 

i, 


Excretion 


Uptake 


Mortality 


Mortality 


»-      PON 


Mortality 


Uptake 


Phytoplankton  (PhyC) 


Uptake 


PhyP^ 'PhyC 


Uptake 


PHOSPHOROUS 


Excretion 


SRP 


n  i  4 


PIP 


Sorption/ 


Desorption 


Zooplankton  (ZooC) 


Mortality 


•H         POP        r* 


Mortality 


Mortality 


Mineralization 


SOP 


Sorption/ 


Desorption 


Mortality 


Excretion 


Figure  4.1  The  connection  between  nitrogen,  phosphorous  and  carbon  cycle 


45 
4.4  Oxygen  Balance 

Dissolved  oxygen  (DO)  refers  to  the  volume  of  oxygen  contained  in  water.  Five  state 
variables  participate  in  the  dissolved  oxygen  balance:  phytoplankton  carbon  (PhyC), 
ammonia  (NH4),  nitrate  (N03),  carbonaceous  biochemical  oxygen  demand  (CBOD),  and 
dissolved  oxygen  (DO).  A  summary  is  illustrated  in  Figure  4.2.  The  methodology  for  the 
analysis  of  dissolved  oxygen  dynamics  in  natural  water,  particularly  in  streams,  rivers,  and 
estuaries  is  reasonably  well-developed  (O'Connor  and  Thomann,  1972).  Dissolved  oxygen 
evolution  depends  on  the  balance  between  production  from  photosynthesis,  consuming  from 
respiration  and  mortality,  and  exchanges  with  the  atmosphere  and  sediment. 

The  main  physical  mechanisms  influencing  DO  concentration  are  horizontal  and 
vertical  dispersion  and  diffusion.  Vertical  diffusion  occurs  across  the  air-sea  interface  as  a 
function  of  wind,  waves,  currents,  and  DO  saturation  rate.  Laboratory  experiments  show  that 
the  bottom  shear  stress  controls  the  dissolved  oxygen  diffusive  layer  thickness  and  the  flux 
at  the  sediment-water  interface  (Steinberger  andHondzo,  1999).  Slow  oxygen  diffusion  rates 
and  high  oxygen  demand  by  sediment  results  in  a  thin  aerobic  layer.  Two  distinct  sediment 
zones  are  created  in  the  sediment  column:  an  aerobic  layer  and  an  anaerobic  layer.  The 
thickness  of  the  aerobic  soil  zone  is  influenced  by  oxygen  concentration  in  the  overlying 
water  column,  and  concentration  of  the  reduced  compounds  in  the  anaerobic  soil  zone.  The 
model  dissolved  oxygen  cycle  includes  the  following  processes 

1 )  Reaeration 

2)  Carbonaceous  oxygen  demand  (CBOD) 

3)  Nitrification 

4)  Sediment  oxygen  demand  (SOD) 


46 


5)  Photosynthesis  and  respiration 


Carbonaceous  Oxygen  Demand 


Zooplankton 

Oxidation 

Mortality 

K^'ZOOC 

Haunt  +  DO 

CBOD 

CBOD 

1 

Mortality 

K^'PHYC 

c 
1 

1 

^ 

f 

r 

Phytoplankton 

dz 


i  Diffusion 


CBOD 


Oxidation 


Denitrification 


DO 


»™»  +DO 


-CBOD        K, 


H  ,»;) 


«„,,+  DO 


-NO  3 


Dissolved  Oxygen 

Air 


Oxidation 


KD — CBOD 

"HiM  +  DO 


Reaeration  K^DOg-DO) 


Phytoplankton 


DO 


Photosynthesis  &  Respiration 


,->• 


Nitrification 


6f^_22_AW4 

14     mHxrT+DO 


1(1.3-0. 


3P.)^-(^+^.)l*PWC»«, 


Sediment  Oxygen  Demand 


KO2+D0 

Water  column 


*  Diffusion 


DO 


Nitrification 


DO 


14      "  H^+DO 


NH4 


Aerobic  Layer 

Figure  4.2  DO  and  CBOD  cycles 
Reaeration 

Reaeration  is  the  process  of  oxygen  change  between  the  atmosphere  and  sea  surface. 
Typically,  dissolved  oxygen  diffuses  into  surface  waters  because  dissolved  oxygen  levels  in 
most  natural  waters  are  below  saturation.  However,  when  water  is  super-saturated  as  a  result 
of  photosynthesis,  dissolved  oxygen  returns  the  atmosphere.  Dissolved  oxygen  saturation 
in  seawater  is  determined  as  function  of  temperature  and  salinity  (APHA,  1985) 


47 


1.575701xl05     6.642308xl07     1.2438x10'°     8.621949x10" 
In  DO. =-139.34411  + —      -  + ~3 =5 


r 

(4.9) 

Sa      I  ,     1.9428x10 

3.1929x10- +  • 


,     1.9428x10     3.8673xl03^ 


1.80655 


T2 


where  DOs  is  equilibrium  oxygen  concentration,  mg/1,  at  standard  pressure 
T  is  temperature,  °K,    °K  =  °C  +  273.150 
Sa  is  salinity,  ppt 
The  reaeration  process  is  modeled  as  the  flux  of  dissolved  oxygen  across  the  water 

surface: 

V^-  =  KAEAs(DOI-DO)  (4-10) 

at 

where  V  and  As  are  volume  and  surface  area  of  the  water  body 

In  case  where  the  air-sea  interface  is  not  constricted,  the  volume  is  V  =  As  ■  Az  .  The 

equation  for  reaeration  can  be  expressed  as 

^.  =  ^AL(D0  -DO)  (4.11) 

dt         Az  V       v 

where  DO  is  dissolved  oxygen  concentration  (mg/1);  KAE  is  reaeration  coefficient  (m/day). 
Many  empirical  formulas  have  been  suggested  for  estimating  reaeration  rate 
coefficient  specially  in  the  river.  Bowie  et  al.  (1985)  have  reviewed  thirty-one  reaeration 
formulas,  and  have  tried  to  evaluate  the  performance  of  the  each  formulas.  Most  formulae 
have  been  developed  based  on  hydraulic  parameters,  most  often  depth  and  velocity.  This 
review  of  stream  reaeration  has  shown  that  no  one  formula  is  best  under  all  conditions,  and 
depending  on  the  data  set  used,  the  range  of  the  reaeration  coefficients  in  the  data  set,  and 
error  measurement  selected,  the  best  formula  may  change.  Among  these  formulas,  the  most 


48 
common  method  of  simulating  reaeration  in  rivers  is  the  O'Connor-Dobbins  formula.  This 
method  has  the  widest  applicability  being  appropriate  for  moderate  to  deep  streams  with 
moderate  low  velocities.  With  approximately  2.09xl0"5  cm2/s  diffusivity  of  oxygen  in 
natural  waters,  the  O'Connor-Dobbins  formula  can  be  expressed  as 

jjO.5 

Ku=3.93—  (4.12) 

For  standing  water,  such  as  lake,  impoundments,  and  wide  estuaries,  wind  becomes 

the  predominant  factor  in  causing  reaeration.  The  oxygen-transfer  coefficient  itself  can  be 

estimated  as  a  function  of  wind  speed  by  a  number  of  formulas.  Chapra  (1997)  compared 

four  common  wind-dependent  reaeration  formulas:  Broeckeret  al.  (1978),  Banks  andHerrera 

(1977),  O'Connor  (1983),  and  Wanninkhof  et  al.  (1991).  The  comparison  shown  in  Figure 

4.3  show  all  these  methods  except  Broeckeret  al.'s  have  similar  reaeration  coefficients  when 

wind  speed  is  less  than  5  m/s.  When  wind  speed  is  greater  than  5  m/s,  Bank  and  Herrera's 

formula  produce  the  middle  range  of  reaeration  coefficient  among  these  three  formulas. 

This  formula  uses  various  wind  dependencies  to  attempt  to  characterize  the  difference 

regimes  that  result  at  air-water  interface  as  wind  velocity  increase  (Banks  1975;  Banks  and 

Hen-era,  1977). 

Kl=0.72SU°w5-03llUw  +  0mi2Ul  (4.13) 

Since  estuary  gas  transfer  can  be  affected  by  both  water  and  wind  velocity,  effort  to 
determine  reaeration  in  estuaries  combines  elements  of  current  and  wind-driven  approaches. 
Thomann  and  Fitzpatrick  (1982)  combined  the  two  approaches  for  estuaries  affected  by  both 
tidal  velocity  and  wind, 


KAE  =  3.93J^  +  0.728£/° 5-  0.3 17£/„ -0.0372£/H2  (4.14) 


49 

where  U0  is  depth  averaged  velocity  (m/s);  H  is  a  depth  (m);  Uw  is  wind  speed  (m/s). 

For  the  Charlotte  Harbor  estuarine  system,  The  aeration  coefficient  is  assumed  to  be 
proportional  to  the  water  velocity,  depth,  and  wind  speed  following  Thomman  and 
Fitzpatrick  (1982). 


10r- 

9  - 
8  - 
7  - 


n 
?     5 


*      4 

3 
2 
1 
0 


Broeckeretal.  (1978)  ,' 


Wanninkhofetal(1991). 


Banks  and 
Herrera(1977) 


4  6 

Uw  (m/s) 


Figure  4.3  Comparison  of  wind-dependent  reaeration  formulas. 
Carbonaceous  Oxygen  Demand 

The  use  of  carbonaceous  oxygen  demand  (CBOD)  as  a  measure  of  the  oxygen- 
demanding  processes  simplified  modeling  efforts  by  aggregating  their  potential  efforts 
(Ambrose  et  al.,  1994).  Oxidation  organic  matter,  nitrification,  non-predatory  mortality  and 
respiration  by  zooplankton  and  phytoplankton  are  nitrogenous-carbonaceous-oxygen- 
demand,  collectively  combined  into  the  state  variable  CBOD. 


50 
The  kinetic  pathway  of  CBOD  is  represented  in  the  source  term  of  the  equation  as 
(Ambrose  et  al.,  1994): 


For  water  column: 

?-CBOD  =  -*- 
dt  dz 


IcBOD  =  -f  [wsCB0D  •  (1  -  fdCB0D )  ■  CBOD]  -  KD  ■        °°        ■  CBOD 
at  dz  tl  r-o^-r  UU 

(4.15) 

--■  —  ■KDN ^ NO,+  Aoc(Kar  PhyC  +  KaZooC) 

4   14  Hno3+DO  V 

For  sediment  column: 

IcBOD  =  +^-[wsCBOD  ■  (1  -  fdCB0D )  ■  CBOD] -  KD  ■         °°        ■  CBOD 

(4.16) 

--■  —  ■KDN ^ NO, 

4   14  Hnu3  +  DO 

where  fdCB0D  corresponds  to  the  fraction  of  the  dissolved  CBOD;  wsCBOD  is  the  settling 
velocity  for  the  particulate  fraction  of  CBOD  ;  KD  is  oxidation  coefficient  which  is  a 
temperature  function;  HCB0D  is  a  half  saturation  constant  for  denitrification;  KDN  is 
denitrification  constant;  Hm3  is  a  half  saturation  rate  for  denitrification;  and  Aoc  is  oxygen- 
carbon  ratio  (g02/gC). 

The  consumption  of  oxygen,  as  a  function  of  water  column  CBOD  decay,  can  be 
expressed  as  CBOD  oxidation; 

—  DO  =  KD — CBOD  (4.17) 

dt  HCB0D  +  DO 

Note  that  non-oxidative  processes  such  as  settling,  denitrification,  and  mortality  do 
not  contribute  to  dissolved  oxygen  depletion,  and  are  not  included  in  the  expression. 

Nitrification 

The  transformation  of  reduced  forms  of  nitrogen  to  more  oxidized  forms 
(nitrification)  consumes  oxygen.  Although  nitrification  is  also  a  nutrient  transformation,  this 
section  addresses  oxygen  consumption.  First  order  kinetics  is  the  most  popular  approach  for 


51 
simulating  nitrification  in  natural  system: 

±DO—£k„— ^—  -NH4  (4.18) 

dt  14     m    Hnil+DO 

where  KNN  is  nitrification  rate  which  is  a  function  of  temperature;  and  Hnil  is  the  half- 
saturation  constant  for  the  bacteria  growth. 

Sediment  Oxygen  Demand 

Sediment  oxygen  demand  is  a  dissolved  oxygen  flux  at  water/sediment  interface  due 
to  the  oxidation  of  organic  matter  in  bottom  sediments.  The  particulate  organic  matters  are 
from  a  source  outside  the  system  such  as  wastewater  particulate  or  leaf  litter  materials,  and 
generated  inside  system  as  occurs  with  plant  growth  in  highly  productive  environments. 
According  to  accumulate  these  particulate  organic  matter  in  the  bottom  sediment,  the 
sediment  oxygen  demand  will  increase  due  to  oxidation  of  the  accumulated  organic  matter 
at  aerobic  sediment  layer.  Figure  4.4  shows  the  mechanism  of  SOD  flux  with  particulate 
organic  matters  in  the  sediment  column.  Particulate  organic  matter  (POM)  is  delivered  to 
the  sediments  by  settling.  Within  the  anaerobic  sediment  layer,  the  organic  carbon,  sulfate, 
nitrate  undergo  reduction  reactions  to  yield  dissolved  methane,  sulfide,  dissolved  ammonium 
nitrogen.  These  reduced  species  (CH4,  H2S,  NH4)  diffuses  upward  to  the  aerobic  layer 
where  these  are  oxidized.  During  these  oxidation  reactions,  SOD  is  generated.  Any  residual 
reduced  species  that  are  not  oxidized  in  the  aerobic  layer  is  diffused  back  into  the  water 
column  where  additional  oxygen  is  consumed  by  oxidation.  Developing  an  SOD  model  is 
quite  complicated  task  since  many  aerobic  and  anaerobic  reactions  in  the  sediments  are 
involved  and  many  chemical  species  are  included  in  mass  balance  equation.  Therefore,  these 
redox  chemistry  with  POM  are  usually  treated  as  a  composite  characteristics  of  the  particular 
system.  Recently,  techniques  have  been  developed  for  investigating  these  factors. 


52 
DiToro  et  al.  (1990)  developed  a  model  of  the  SOD  process  in  a  mechanistic  fashion 
using  the  square-root  relationship  of  SOD  to  sediment  oxygen  carbon  content.  Using  similar 
analysis  as  applied  to  carbon,  they  also  evaluate  the  effect  of  nitrification  on  SOD.  In  this 
model,  carbon  and  nitrogen  diagenesis  are  assumed  to  occur  at  uniform  rates  in  a 
homogeneous  layer  of  the  sediment  of  constant  depth  (active  layer).  The  sediment  oxygen 
demand  and  sediment  fluxes  are  calculated  by  the  concentrations  of  particulate  organic 
carbonaceous  material  and  of  particulate  organic  nitrogenous  material  in  this  active  layer. 
The  more  detail  description  of  sediment  flux  model  developed  by  DiToro  and  Fitzpatrick 
(1993)  is  presented  in  Appendix  F.  Their  framework  has  been  applied  to  the  Chesapeake 
Bay  (Cerco  and  Cole,  1994;  DiToro  and  Fitzpatrick,  1993)  with  sediment  flux  model  which 
include  ammonia  and  nitrate  flux  and  the  sulfide,  oxygen,  phosphorous  and  silica  flux. 

A  calculation  of  the  detailed  redox  chemistry  of  the  sediment  interstitial  water  is 
required  for  a  detail  understanding  of  the  situation  for  each  reduced  species  and  chemical 
parameters  for  redox  reaction  such  as  pH,  Eh  in  the  system.  In  the  absence  of  data  for  CH4, 
H2S,  Iron,  Methane,  and  C02,  and  detail  understanding  of  redox  chemistry  of  the  system, 
applying  this  model  could  create  more  uncertainty  than  the  simple  empirical  formula  based 
on  the  measurement  technique. 

The  process  of  oxygen  demand  in  the  sediment/water  interface  is  usually  referred  to 
as  sediment  oxygen  demand  (SOD)  because  of  the  typical  mode  of  measurement:  enclosing 
the  sediments  in  the  chamber  and  measuring  the  change  in  the  dissolved  oxygen 
concentration  at  several  time  increments.  The  major  factors  affecting  SOD  are:  temperature, 
oxygen  concentration  at  the  sediment  water  interface,  organic  and  physical  characteristics  of 
the  sediment,  and  current  velocity  over  the  sediments  (Bowie,  1985). 


Water  Column 


SL 


53 


02 


Aerobic  Layer 


Oxidation 


CH4,  NH4,  H2S 


Anaerobic  Layer 


Diffusion 


Reduction 


^     CH4,  NH4,  H2S 


Production  of 
CH4,  NH4,  H2S 


Figure  4.4  The  relationship  between  POM  flux  and  SOD  flux  related  in  the  oxidation  and 
reduction  of  organic  matter  in  sediment  column. 


54 


Typical  values  of  SOD  are  listed  in  Table  4.1  (Chapra,  1997).  In  general,  values  from 
about  1  to  10  g02/m2-day  are  considered  indicative  of  enriched  sediments.  According  to  this 
table,  SOD  at  a  mud  bottom  is  higher  than  SOD  at  a  sandy  bottom.  In  Charlotte  Harbor 
estuarine  system,  the  sediment  type  near  upper  Charlotte  Harbor  is  finer  and  includes  more 
clay  and  mud  than  those  for  the  other  study  area.  The  finer  sediment  which  has  clay,  silt,  and 
mud  would  have  higher  SOD  as  more  organically  rich  sediments.  From  this  assumption, 
SOD  can  be  applied  as  a  function  of  sediment  bottom  type  (sediment  size). 
Table  4.1  Average  Values  of  Oxygen  Uptake  Rates  of  Bottom  (Chapra,  1997) 


SOD  rate  at  20°C  (g02/m2-day  ) 


Bottom  type  and  Location 


Average  Value 

Range 

7 

4 

2-10 

1.5 

1-2 

1.5 

1-2 

0.5 

0.2-1 

0.07 

0.05-1 

Sphaerolitus  (10  g-dry  wt — 2) 

Municipal  sewage  sludge: 
Outfall  vicinity 
Downstream  of  outfall 

Estuarine  mud 

sandy  bottom 

Mineral  soil 


The  effect  of  temperature  and  sediment  type  on  SOD  can  be  represented  by 


Sb(T)  =  Sb,20'ST  '& 


7-20 


(4.19) 


where  SB20  is  an  areal  SOD  rate  at  20  °C  (g02/nr-day  )  which  is  user  defined  value;  ST  is 
a  fractional  coefficient  for  sediment  type.  (When  sediment  type  <2,  ST=1,  when  sediment 
type  >3,  ST=0.5) ;  0  is  temperature  coefficient.  Zison  et  al.  (1978)  have  reported  a  range  of 
1.04  to  1.13  for  0.  A  value  1.065  is  commonly  employed  and  is  used  in  this  study. 

Oxygen  is  another  factor  that  affects  sediment  oxygen  demand.  Sediment  oxygen 
consumption  is  reduced  as  oxygen  concentration  in  the  overlying  water  decrease.  Lam  et  al. 
(1984)  use  a  Michealis-Menten  relationship  to  represent  the  dependence,  by  a  saturation 
relationship, 


55 
SB(DO)=       D°SB(T)  (4.20) 

KSOD+UU 

where  KSOD  is  half  saturation  rate  for  SOD.  Lam  et  al.  (1984)  have  suggested  a  value  for  KS0D 
of  1.4  mg/1. 

The  decay  of  substrate  is  assumed  to  balance  continued  settling  resulting  in  a  steady- 
state  sediment  concentration  of  oxygen  demand  substrate.  According  to  this  assumption,  the 
kinetic  equation  for  sediment  oxygen  demand  is  (Ambrose  et  al.,  1994): 

dDO__SOD  (421) 

dt  H 

where  H  is  water  depth  (m);  and  SOD  is  sediment  oxygen  demand  (as  measured),  g02/m2- 
day.  SOD  can  be  calculated  as  a  function  of  temperature,  dissolved  oxygen  at  the  water- 
sediment  interface,  and  sediment  bottom  type  based  on  characteristics  of  SOD  measurement. 

Exchange  of  material  between  the  water  column  and  benthic  sediment  is  an  important 
component  of  the  eutrophication  process.  Sediment  oxygen  demand  may  comprise  a 
substantial  fraction  of  total  system  oxygen  consumption  (Cerco  and  Cole,  1995). 

Oxygen  consumption  in  the  sediments  depends  upon  water-column  temperature  and 
oxygen  availability,  and  sediment  type.  As  temperature  increases,  respiration  in  the  sediment 
increases.  Sediment  oxygen  consumption  is  reduced  as  oxygen  concentration  in  the 
overlying  water  decreases.  Therefore,  the  kinetic  equation  for  sediment  oxygen  consumption 
(SOC)  in  sediment  column  can  be  represented  as  (Cerco  and  Cole,  1995): 


dDOsed         SOC  _      1    s        QT^ DO, 

dt  H  H      B-2°  KS0D+1 


ML  -  -^1  =  -  J-  .  5  .  flr-20  .    _      f^Mfc (422) 


where  SB20  is  a  function  of  sediment  bottom  type  (from  table  4.1). 

The  processes  that  create  sediment  oxygen  demand  are  little  affected  by  the 


56 

concentration  of  oxygen  in  the  overlying  water.  When  oxygen  is  unavailable  to  fulfill 
sediment  oxygen  demand,  the  demand  is  exported  to  the  water  column.  The  exported 
demand  may  be  in  the  form  of  reduced  iron,  manganese,  methane,  or  sulfide,  which  are 
represented  in  the  model  as  sediment  oxygen  demand  and  provides  a  function  which 
computes  additional  release  as  oxygen  consumption  in  the  sediment  is  restrained  (Cerco  and 
Cole,  1995). 

d£O_SOD=_l_T.20t_K^_ 

dt  H  H      Bao  KSOD  +  DO 

The  relationship  between  sediment  oxygen  consumption  and  sediment  oxygen 

demand  is  represented  in  Figure  4.5.  The  SOD  is  negligible  when  DO  much  higher  than 

KSOD.    When  dissolved  oxygen  is  absent  from  the  water  column,  the  maximum  oxygen 

demand  is  released  to  the  water  as  sediment  oxygen  demand. 


57 


SOC  (Sediment  Oxygen  Consumption)  at  sediment  column 
SOD  (Sediment  Oxygen  Demand)  at  water  column 

where    SB  (T=  20°C)  =  2  g/m2/day 
KS0D      =  2  g/m3 


X 

2-0.5 


■1 


■1.5 

-2 


Dissolved  Oxygen  (g/m  ) 


Figure  4.5  Effect  of  dissolved  oxygen  on  sediment  oxygen  consumption  and  SOD  release 


58 
Photosynthesis  and  Respiration 

The  photosynthesis  and  respiration  of  phytoplankton  can  add  and  deplete  significant 
quantities  of  oxygen  from  natural  systems.  The  produced  oxygen  concentration  by 
photosynthesis  depends  on  the  form  of  the  nitrogen  species  accessed  for  phytoplankton 
growth.  One  mole  carbon  dioxide  can  produce  one  mole  oxygen  when  ammonium  is  the 
nitrogen  source,  while  one  mole  carbon  dioxide  produces  1.3  moles  oxygen  when  nitrate  is 
the  nitrogen  source,  according  to  Morel's  equation  (1983) 

106Ca  +\6NH:  +  H.PO;  +  106ff2O  -»  protoplasm  +  l06O2  +  15H+ 

(4.24) 

\06CO2+\6NO~  +H2PO;  +\22H20  +  UH^  ->  protoplasm  +  13802 

The  simple  representation  of  the  respiration  process  can  be  used  to  determine  how 
much  oxygen  would  be  consumed  in  the  decomposition  of  a  unit  mass  of  organic  carbon, 

6C02  +  6H20  ^  C6Hl206  +  602  (4.25) 

The  dissolved  oxygen-to-carbon  ratio  in  respiration  can  be  calculate  from  this 
equation. 

Aoc  =  (^-  =  2.67gO/gC  (4.26) 

oc      6(12) 

The  equation  that  describes  photosynthesis  and  respiration  on  dissolved  oxygen  is: 
^  =  [(1.3 - 0.3 •  PH)jUa  ~ K„  -  KM ]  •  Aoc  ■  PhyC  (4.27) 

where  Pn  is  nitrogen  preference  coefficient;  //a  is  phytoplankton  growth  rate,  which  is  a 
function  of  the  intensity  of  light,  the  availability  of  nutrients,  and  the  ambient  temperature; 
and  Kas  and  K^  are  respiration  and  mortality,  which  are  functions  of  temperature. 

The  mass  balance  equation  for  dissolved  oxygen  is  written  by  combining  all  oxygen 
transformation  processes. 


59 


For  water  column: 


^  =  +^{D0  -DO)-KD 52 CBOD 

dl  teK      '  '  HCBm  +  DO 

14     m    HNIT+DO  Az 

+[(l3-03.Pn)jua-K(U-Kax}Aoc-PhyC 

which  include  reaeration,  oxidation  by  CBOD,  nitrification,  sediment  oxygen  demand,  and 

photosynthesis  and  respiration  terms. 
Fore  sediment  column: 

dDO  DO  „   _     64  „  DO         XTr,      SOC 

=  -KD CBOD KNN NH4 (4.29) 

&  HCBOD+DO  14     NN    HNIT  +  DO  Az 

which  include  oxidation  by  CBOD,  nitrification,  and  sediment  oxygen  consumption. 
4.5  Effects  of  Temperature  and  Light  Intensity  on  Water  Quality  Processes 

Temperature  and  light  intensity  are  the  most  important  parameters  for  transformation 
processes.  To  achieve  a  better  understanding  of  water  quality  processes,  it  is  necessary  to 
improve  the  spatial  and  temporal  variation  of  these  parameters. 
4.5.1  Temperature 

In  the  nutrient  cycle,  almost  all  the  reaction  parameters  are  affected  by  temperature, 
such  as  zooplankton  and  phytoplankton  growth,  respiration  and  mortality,  nitrification, 
denitrification,  NH3  stability,  mineralization,  oxidation,  sediment  oxygen  demand,  and 
sorption/desorption  reactions.  The  effect  of  temperature  on  reaction  rates  can  be  explained 
by  the  Van't  Hoff-Arrhenius  equation,  as  follows: 

^^1  =  ^L  (4.30) 

dt         RT2 

where  K  is  reaction  rate  at  temperature  T,  AH  is  the  amount  of  heat  required  to  bring  the 

molecules  of  the  reactant  to  the  energy  state  required  for  the  reaction,  and  R  is  universal  gas 


60 


constant. 


Integrating  Equation  (4.44)  from  temperature  T,  to  T2  gives: 


K2 

— -  =  exp 


AH 
RT{T2 


(ra-rt) 


(4.31) 


where  K,  and  K2  are  reaction  rate  at  temperature  T,  and  T2 ,  respectively.  The  temperature 


adjustment  function  ( 6  =  exp 


A// 
RTJ2 


)  is  almost  constant  at  the  temperature  range  of 


interest  (0°  ~  30°C),  ranging  from  1.01  to  1.2.  This  equation  can  be  rearranged  into  a  more 
useful  form  as: 


K(T)  =  K(Tref)-0{T~T"/> 


(4.32) 


where  K(T)  is  the  reaction  coefficient  at  temperature  T,  such  as  /ua  (phytoplankton  growth 
rate),  K^  (phytoplankton  respiration  rate),  Km  (phytoplankton  mortality  rate),//,  (zooplankton 
growth  rate),  Ka  (zooplankton  respiration  rate),  K^  (zooplankton  mortality  rate),  KM 
(ammonia  instability),  Kom  (ammonification  rate),  KNN  (nitrification  rate),  dun 
(sorption/desorption  rate  for  organic  nitrogen),  dm  (sorption/desorption  rate  for  inorganic 
nitrogen),  K0PM  (mineralization  rate),  dop  (sorption/desorption  rate  for  organic  phosphorous), 
dap  (sorption/desorption  rate  for  inorganic  phosphorous),  KD  (oxidation  rate),or  SOD 
(sediment  oxygen  demand).  Each  rate  term  has  a  unique  temperature  adjustment  function. 
Most  models,  which  use  exponential  temperature  functions,  assume  a  reference 
temperature  of  20  °C  (Chen  and  Orlob,  1975;  Thomann  and  Frizpatrick,  1982).  Eppley 
(1972)  showed  that  an  exponential  relationship  describes  the  envelope  curve  of  growth  rate 
versus  temperature  data.  The  determination  was  made  with  a  large  number  of  studies,  with 


61 
many  different  species. 

4.5.1  Light  intensity 

Light  intensity  affects  the  photosynthesis  process  and  thus  the  phytoplankton  growth 
rate.  The  effects  of  light  intensity  on  nutrient  cycling  is  often  modeled  by  a  light  intensity 
limiting  function  as  follow  (Steele,  1974) 


/(/)  =  y-exp 


1-1 


(4.33) 


where  /  is  the  light  intensity,  and  Z,  is  the  optimum  light  intensity  for  phytoplankton  growth 
.  According  to  the  Lambert-Beer  equation,  the  light  intensity  over  the  water  depth  is: 

I(z)  =  I0-exp[-Kd(PAR)-z]  (4.34) 

where  I(z)  is  the  light  intensity  at  depth  z,  and  I0  is  the  light  intensity  at  the  water  surface. 
KJPAR)  is  a  function  of  suspended  sediment  concentration,  algae  concentration,  and  color. 
This  value  was  calculated  by  the  light  attenuation  model,  which  will  be  discussed  in  the  next 
section. 

4.6  Light  Attenuation  Model 

One  of  the  most  important  variables  controlling  phytoplankton  photosynthesis  is 
"photosynthetically  active  radiation"  (PAR),  or  light,  in  the  range  of  wavelengths  from  400- 
700nm,  which  provides  the  predominant  source  of  energy  for  autotrophic  organisms  (Day 
et  al.,  1989).  Absorption  and  scattering  of  light  by  water  and  dissolved  and  suspended 
matter  determine  the  quantity  and  spectral  quality  of  light  at  a  given  depth  (Jerlov,  1976; 
Prieur  and  Sathyendranath,  1981),  which  in  tern  affect  the  photosynthesis  of  aquatic  plants. 

One  way  to  develop  a  light  attenuation  model  is  to  find  a  simple  regression 
relationship  between  Kd  (PAR),  calculated  from  light  measurements,  and  water  quality 


62 
measurements  collected  at  the  same  instant  as  the  Kd  (PAR)  vales.  This  type  of  empirical 
model  is  a  simple  way  to  relate  light  attenuation  to  water  quality,  at  a  certain  time.  This 
method  was  used  by  Mcpherson  and  Miller  (1994)  in  Tampa  Bay  and  Charlotte  Harbor.  A 
physics-based  light  attenuation  model  was  developed  as  part  of  the  CH3D-EVIS  (Sheng  et  al. 
2001c,  Christian  and  Sheng,  2003)  and  successfully  applied  to  the  Indian  River  Lagoon.  This 
light  model  is  adopted  for  the  Charlotte  Harbor  study. 

In  the  light  model,  the  vertical  light  attenuation  coefficient  (Kd)  is  a  function  of  solar 
zenith  angle  (yu0),  scattering  (b)  and  total  absorption  (a,)  (Kirk,  1984) 

1    r     ,  ,        N  -|l/2 

Mo  J  (4.35) 

and     G(p0)  =  grfi0-g2 

with  g,  =0.473  and  g2=0.218  determined  for  the  mid  point  of  euphotic  zone  (Kirk,  1984). 
The  scattering  coefficient,  b,  is  determined  only  using  particles  since  scattering  due 
to  particles  is  much  greater  than  scattering  due  to  water  (Gallegos,  1994).  Scattering  can  be 
described  as  a  function  of  turbidity  (Morel  and  Gentili,  1991)  as  follow: 

b(A)  =  (550/ A)  ■  [Turbidity]  (4.36) 

Total  absorption  (a,)  is  partitioned  into  attributes  of  water  (aw),  phytoplankton  (aph), 
dissolved  color  (ad(),  and  detritus  {ad). 

a,  =  «»•  +  aPh  +  adc  +  ad  (4.37) 

The  absorption  of  water  (a  J  can  be  determined  from  literature  values  (Smith  and 
Baker,  1981)  with  1  nm  linearly  interpolated  from  5  nm  found  in  the  literature.  Chlorophyll- 
specific  absorption  (aph)  is  calculated  for  the  model  from  the  linear  relationship  between  the 
maximum  absorption  and  analytical  chlorophyll_a  concentration  (Dixon  and  Gary,  1999): 


63 

anh  C^)  =  \aoh )       '  formalized  _  Spectra(A) 

MAX  (4.38) 

fa  .  J       =0.0209 -(Chlorophyll  a,    corrected) 

where  (aph)MAX  is  the  maximum  absorption  of  chlorophyll-a.  To  calculate  normalized_spectra 
(X),  individual  spectra  are  normalized  to  the  maximum  absorption  (437  -  440  nm),  and 
averaged  for  all  samples  for  an  overall  normalized  spectra. 

The  absorption  by  dissolved  color,  adc,  for  each  wavelength  in  the  visible  spectrum 
can  be  found  using  a  negative  exponential  function  (Bricaud  et  al.,  1981) 

adlU)  =  8uo  -«p[-**  (A -440)]  (4.39) 

where  g440  is  the  absorption  by  dissolved  color  at  440  nm  and  sdc  is  spectral  slope.  Dixon 
and  Gary  (1999)  calculated  empirical  absorption  at  440  nm  and  spectral  slope  as  a  function 
of  color  (in  PCU)  at  Charlotte  Harbor  in  the  form: 

g44()  =  0.0667  ■  f color] ,         n  =  129,  r  =  0.9329 

(4.40) 
sJc  =  0.00003  •  [color]  -  0.0178,         n  =  129,  r  =  0.5 1 1 1 

Absorption  due  to  organic  and  mineral  detritus  is  represented  as  a  function  of 
turbidity  (Gallegos,  1994): 

ad  =  Gd  (A)  •  [Turbidity]  (4.41) 

where  turbidity  is  in  NTUs  and  od(A)  is  the  wavelength  specific  absorption  cross  section  of 
turbidity  as  calculated  in: 

<?d  U)  =  <*bi  +  ^oo  •  exP  [~sd  (A  ~  40°)]  (4-42) 

in  which  abl  is  the  longwave  absorption  cross  section,  a400  is  the  maximum  detritus 
absorption  at  400  nm,  and  sd  is  exponential  slope. 

The  total  absorption  and  scattering  are  used  in  Equation  (4.49)  to  calculate  the 


64 
vertical  attenuation  coefficient,  KJA),  for  each  wavelength  depending  on  color, 
chlorophylls  ,  and  turbidity.  This  KJA)  value  and  incident  irradiation  EJA)  can  be  used  in 
the  equation  for  calculating  irradiance  EZ(A)  at  the  reference  depth,  zr: 

Ez(A)  =  E0(A)-exp[-Kd(A)-zr]  (4.43) 

The  spectrum  of  incident  sunlight  data  from  table  F-200  in  Weast  (1977)  is  used  for 
incident  spectral  information,  EJA),  as  in  Gallegos's  work  (1994) .  These  data  are  shown 
in  Table  4.2. 


Table  4.2  The  spectrum  of  incident  sun 


X  (nm) 


400 

405 

410 

415 

420 

425 

430 

435 

440 

445 

450 

455 

460 

465 

470 

475 

480 

485 

490 

495 

500 


X  (nm) 


4.780 

5.568 

505 

6.003 

510 

6.052 

515 

6.135 

520 

6.017 

525 

5.893 

530 

5.940 

535 

6.659 

540 

7.152 

545 

7.548 

550 

7.826 

555 

7.947 

560 

7.963 

565 

7.990 

570 

8.119 

575 

8.324 

580 

8.014 

585 

7.990 

590 

8.113 

595 

8.119 

600 

ight  data  (Gallegos,  1994) 


X  (nm) 


8.108 

505 

8.026 

510 

7.894 

515 

7.970 

520 

8.130 

525 

8.163 

530 

8.133 

535 

8.051 

540 

7.993 

545 

7.993 

550 

7.982 

555 

7.937 

560 

8.055 

565 

8.160 

570 

8.265 

575 

8.318 

580 

8.375 

585 

8.387 

590 

8.369 

595 

8.359 

600 

8.332 
8.340 
8.323 
8.305 
8.289 
8.271 
8.267 
8.263 
8.239 
8.213 
8.207 
8.201 
8.180 
8.157 
8.136 
8.114 
8.106 
8.089 
8.052 
8.013 


EJA)  and  E,(A)  are  integrated  over  the  visible  spectrum  to  get  PAR0  for  the  incident 


PAR  and  PARZ  for  the  calculated  PAR  at  the  reference  depth.  The  spectrally  sensitive 


65 

attenuation  coefficient  Kd{PAR)  is  calculated  from  these  integrated  values  by  using  the 

rearranged  form  of  the  Lambert-Beer  equation: 


~  1       < 

Kd(PAR)  =  —\n 

Z. 


PARZ 


V  PARo  j 


(4.44) 


This  Kj(PAR)  will  be  used  in  the  model  to  calculate  light  levels  throughout  the 

water  column  as  a  function  of  measured  incident  light  intensity.  Dixon  and  Gray  (1999) 
calibrated  the  light  attenuation  model  with  measured  data  and  applied  the  model  to  determine 
the  light  requirements  for  seagrasses  of  the  Charlotte  Harbor  estuarine  system.  The  results 
of  the  optical  model  are  very  good,  with  mean  percentage  agreement  between  modeled  and 
observed  Ktl  of  1 10%.  Using  water  quality  data  from  all  stations  as  inputs,  Dixon  and  Gray 
(1999)  determined  that  chlorophyll,  color,  and  turbidity  account  for  4%,  66%,  and  31%  of 
water  column  light  attenuation,  respectively.  The  maximum  annual  average  chlorophyll 
contribution  is  6%,  with  an  individual  maximum  of  18%  during  a  phytoplankton  bloom. 
Color  dominated  water  column  attenuation,  ranging  from  40%  to  78%.  Stations  in  the  lower 
Harbor  and  southern  sites  showed  increased  attenuation  due  to  turbidity,  compared  to  the 
upper  Harbor  site  (Dixon  and  Gary,  1999).  For  the  southern  sites,  a  much  larger  portion  of 
light  attenuation  is  produced  by  turbidity,  up  to  55%  for  the  station  near  Captiva  pass. 

To  provide  a  light  attenuation  component  which  can  be  coupled  with  water  quality, 
hydrodynamic,  and  sediment  model  components  in  the  Charlotte  Harbor  estuarine  system. 
A  stand  alone  light  model  needs  to  be  developed  and  calibrated  with  measured  data.  The 
stand  alone  light  model  has  been  calibrated  by  finding  best  fit  between  simulated  and 
measured  light  attenuation  with  various  coefficient  sets. 

For  calibration  of  the  light  model,  aM,  a400,  sd,  and  s  are  allowed  to  vary  within 


66 
certain  ranges.  The  predicted  kJPAR)  values  for  the  stand  alone  light  model  are  compared 
to  the  corresponding  kJPAR)  values  from  data  provided  by  the  SFWMD.  If  the  predicted 
values  do  not  match  data  well,  then  attempts  are  made  to  try  to  find  model  coefficients  to 
produce  better  fit.  In  this  case,  the  model  is  run  and  the  coefficients  are  varied  between 
maximum  and  minimum  literature  values  specified  in  Table4.3  (Christian  and  Sheng,  2003). 
More  than  one  thousand  runs  of  light  model  simulations  were  conducted  with  this  data  set. 
The  RMS  errors  were  calculated  for  each  run  to  test  model  performance  of  each  set  of  light 
model  coefficients.  The  root  mean  square  error  is  an  indication  of  the  average  discrepancy 
between  observations  and  model  results.  In  addition  to  root  mean  square  error,  model 
calibration  was  assessed  via  plots  of  model  output  and  observation.  Scatter  plots  of  model 
output  and  observed  data  provide  an  indication  of  overall  model  performance.  The  best 
RMS  errors  for  each  of  the  tests  is  0.61 1 1  — '  with  best  fit  coefficients  shown  in  Table  4.4 
Dixon  and  Gray  (1999)  found  the  coefficients  in  Table  4.5  provided  the  best  fit  to  all  the  data 
they  examined.  They  applied  a400  and  sy  as  a  function  of  color.  The  RMS  error  with  this 
coefficient  is  0.65 11m"1.  Even  though,  they  reduced  adjustable  coefficients  with  relationship 
of  a400  andsy  with  color,  the  RMS  error  is  greater  than  that  for  adjusting  all  four  coefficients. 
In  this  study,  the  best  fit  coefficients  were  used  for  simulating  light  attenuation.  Figure  4.6 
contains  calibration  period  scattering  plots  with  best  fit  of  coefficient  sets.  The  location  of 
circles  indicate  the  correlation  between  model  predictions  and  observed  data.  A  perfect 
match  between  model  and  observed  data  is  indicated  by  the  diagonal  line  on  each  graph. 
Circle  above  the  line  shows  over  prediction,  while  circles  below  the  line  indicate  that  under 
prediction. 


67 


Table  4.3  Coefficient  ranges  for  use  in  stand  along  light  model. 


Coefficients 


Minimum  value 


o,. 


o_ 


41)11 


0.004 
0.255 
0.011 
0.011 


Maximum  value 


0.090 
0.600 
0.019 
0.019 


Table  4.4  Best  fit  light  model  coefficients  for  Charlotte  Harbor  estuarine  system. 


Coefficient 


<Ju 


aA 


401) 


Value 


0.01  m-'NTTJ-1 

0.59  m'NTU'' 

0.014  nm1 

0.013  nm'1 


Table  4.5  Dixon  and  Gray's  model  coefficients  for  the  Charlotte  Harbor  estuarine  system. 

Coefficient 

Value 

°u 

0.064  m'NTU-1 

°400 

0.0014xcolor+0.0731  m'NTU'1 

*i 

0.01125  nm1 

Sv 

0.00003xcolor+0.0178  nm ' 

68 


T3 
0 


■o 


cr 
< 

Q. 


5  - 


4  - 


3  - 


2  - 


1  - 


0 


•         PARPS  model 

• 

* 

—                                                                                                                                                                                         • 

0 

_                                                                                                                                                                                                                                                                                                                                       0 

0 

0 
_                                                                                                                                                                                                                                                                                                                       0 

0 
0 

-                                                                                                                                                                                                                                                                                                        0 

^m                                                                                                                                                                                                               0 

0 
0 

—                                                                                                                                                                                                        0 

0 

0 
0 
0 
0 

0 
_                                                                                                                                                                                                             0 

• 

0 

0 

• 

— ■                                                                                                                                                                               0 

0 

• 

•    / 

0 
_                                                                                                                                                                                                        # 

0 
0 
0 

•      •  •+'  • 

•*''•  •••              • 

•       %T\0      •                                    # 

• 
_               * 

0 

'**'    I    I    I   I  J    1   1    I   I   I    I    I    I    1   I    I    I    I    1    I    I    I    I   1   I    I    I    I 

2  3  4 

KpAR  observed 


Figure  4.6  The  scatter  plots  for  Kd(PAR)  during  calibration  period  with  best  fit  coefficients. 


69 

The  water  quality  and  sediment  model  outputs  needed  for  the  light  model  are 

turbidity  in  NTUs,  chlorophyll_a  concentration  in  Hg/L,  and  color  in  Pt.  Units.  To  calculate 

solar  angle,  this  model  needs  the  latitude  and  simulation  day  and  time.  Because  the  sediment 

model  simulates  the  total  suspended  solids  (TSS),  it  is  necessary  to  develop  a  regression 

between  TSS  (mg/L)  and  turbidity  (NTU)  as  follow: 

Turbidity  =  0.2677  x  TSS  +  0.9665  for  1996  data 

Turbidity  =  -0.0 1 53  x  TSS  +  2.602  for  2000  data 

The  stand  alone  light  model  was  tested,  calibrated,  and  then  coupled  with  the  models 
of  hydrodynamics,  sediment  and  water  quality. 

4.7  Model  Parameters  and  Calibration  Procedures 

Reaction  terms  described  in  the  previous  section  contain  many  model  parameters 
which  must  be  determined  before  using  the  model.  The  determination  of  model  parameters 
is  generally  very  difficult  because  they  depend  on  many  physical  and  biochemical  factors, 
such  as  the  location  of  the  estuary,  temperature  and  tidal  variation,  and  point  or  non-point 
source  loadings  of  nutrients  and  other  chemical  materials.  In  practice,  parameters  are 
selected  from  a  range  of  feasible  values,  tested  in  the  model,  and  adjusted  until  an  optimal 
agreement  between  simulated  and  measured  values  is  obtained.  Two  ways  to  determine 
feasible  ranges  of  model  parameters  are  field  observation  and  laboratory  experimentation. 
When  field  observations  and  laboratory  experimentation  are  not  available,  the  feasible  ranges 
are  obtained  from  literature  or  previous  modeling  studies.  A  list  of  the  important  kinetic 
parameters  and  their  literature  values  are  given  in  Table  4.6  ~  4.11. 


70 


Table  4.6  Temperature  adjustment  functions  for  water  quality  parameters 


Paramete 

r 

description 

Unit 

literature 
value 

Reference 

(»*) 

temperature  function  for 
phytoplankton  growth 

- 

1.01  ~  1.2 
1.09 
1.08 

Di  Toro  et  al.  (1980) 
Pribble  et  al.(1997) 
Sheng  etal.  (2001) 

K») 

temperature  function  for 
phytoplankton 
respiration  and  mortality 

- 

1.045 

1.05 

1.08 

Ambrose  (1991) 
Pribble  etal.(1997) 
Sheng  et  al.  (2001) 

(»«) 

temperature  function  for 
zooplankton  growth, 
respiration  and  mortality 

- 

1.01  -  1.2 
1.04 

Di  Toro  etal.  (1980) 
Sheng  et  al.  (2001) 

\"oNM  ) 

temperature  function  for 
ammonification 

- 

1.0-  1.04 
1.07 
1.02 

Bowie  et  al.  (1985) 
Pribble  etal.(1997) 
Sheng  et  al.  (2001) 

ifim) 

temperature  function  for 
nitrification 

- 

1.02-1.08 
1.08 
1.08 

Bowie  et  al.  (1985) 
Pribble  et  al.(1997) 
Sheng  et  al.  (2001) 

(<V) 

temperature  function  for 
denitrification 

- 

1.02-  1.09 

1.04 
1.045 

Bowie  etal.  (1985) 
Pribble  et  al.(1997) 
Sheng  et  al.  (2001) 

M 

temperature  function  for 
ammonia  instability 

- 

1.08 

Sheng  etal.  (2001) 

(0OPM  ) 

temperature  function  for 
mineralization 

- 

1.08 

Sheng  etal.  (2001) 

{&s/d) 

temperature  function  for 
sorpti  on/desorpti  on 

- 

1.08 

Sheng  etal.  (2001) 

M 

temperature  function  for 
oxidation 

- 

1.02-1.15 
1.08 

Bowie  et  al.  (1985) 
Sheng  etal.  (2001) 

(8 SOD  ) 

temperature  function  for 
sediment  oxygen 
demand 

- 

1.045 
1.08 

Bowie  et  al.  (1985) 
Sheng  et  al.  (2001) 

71 


Table  4.7  Water  quality  parameters  related  to  conversion  rate 

Paramete 

description 

Unit 

literature 

Reference 

r 

value 

Ajc 

Phytoplankton  Carbon 

go2/gc 

2.67 

Ambrose(1991) 

/  Oxygen  rate 

2.67 
2.67 

Cerco  and  Thomas 
(1995) 
Shengetal.  (2001) 

ChlaC 

Phytoplankton  Carbon 

gC/gChl 

10-  112 

Bowie  et  al.  (1985) 

/  Chlorophyll_a  rate 

a 

100 

60 

50 

Pribble  et  al.(1997) 
Cerco  and  Thomas 
(1995) 
Shengetal.  (2001) 

A:/V 

Phytoplankton  Carbon 

gN/gC 

0.05  -0.43 

Jorgensen  (1976) 

/  Nitrogen  rate 

0.15 
0.167 
0.15 

Pribble  et  al.(1997) 

Cerco  and  Thomas 

(1995) 

Sheng  et  al.  (2001) 

\p 

Phytoplankton  Carbon 

gP/gC 

0.005-0.03 

Jorgensen  (1976) 

/  Phosphorous  rate 

0.027 
0.025 

Cerco  and  Thomas 
(1995) 
Shengetal.  (2001) 

Table  4.8  Water  quality  parameters  related  to  phytoplankton  and  zooplankton 

Paramete 
r 

description 

Unit 

literature 

value 

Reference 

ta>L 

maximum  phytoplankton 
growth  rate 

1/day 

0.2-8 
2.25-2.5 
1.06-2.68 

Bowie  et  al.  (1985) 
Cerco  and  Thomas 
(1995) 
Shengetal.  (2001) 

Hn 

Nitrogen  half  saturation 
rate  for  phytoplankton 
uptake 

ng/i 

1.5-400 
0.5 

1 
10 

Bowie  etal.  (1985) 

Pribble  et  al.(1997) 

Cerco  and  Thomas 

(1995) 

Sheng  et  al.  (2001) 

HP 

Phosphorous  half 
saturation  rate  for 
phytoplankton  uptake 

Mfl 

1.  -  105 
1 

1 

2-4 

Bowie  etal.  (1985) 
Pribble  et  al.(  1997) 
Cerco  and  Thomas 
(1995) 
Shengetal.  (2001) 

72 


opt 

optimum  light  intensity 
for  phytoplankton 
growth 

ly/day 

225-600 
300 

Canale  et  al.  (1976) 
Sheng  et  al.  (2001) 

KM 

phytoplankton 
respiration  rate 

1/day 

0.02-  0.24 
0.03-  0.09 
0.03-  0.05 

Jorgenson  (1976) 

Cerco  and  Thomas 

(1995) 

Sheng  etal.  (2001) 

Kus 

phytoplankton  non- 
predator  mortality 

1/day 

0.01-0.22 
0.03-  0.09 
0.02-  0.06 

Jorgenson  (1976) 

Cerco  and  Thomas 

(1995) 

Sheng  etal.  (2001) 

wsP*y 

phytoplankton  settling 
rate 

m/da 

y 

0.  -  3. 

0.  -  0.25 

0.05-0.1 

Bowie  etal.  (1985) 

Cerco  and  Thomas 

(1995) 

Sheng  et  al.  (2001) 

HPky 

phytoplankton  half 
saturation  rate  for 
zooplankton  uptake 

ng/i 

200-2000 
800-1200 

Bowie  et  al.  (1985) 
Sheng  et  al.  (2001) 

TrSP»y 

phytoplankton  threshold 
for  zooplankton  uptake 

ran 

1-200 
200 

Bowie  et  al.  (1985) 
Sheng  et  al.  (2001) 

V    *  /max 

maximum  zooplankton 
growth  rate 

1/day 

0.15-0.5 
0.18-0.2 

Bowie  et  al.  (1985) 
Sheng  et  al.  (2001) 

K„ 

zooplankton  respiration 
rate 

1/day 

0.003-0.07 

5 

0.01 

Bowie  etal.  (1985) 
Sheng  et  al.  (2001) 

Ka 

zooplankton  non- 
predator  mortality 

1/day 

0.001-0.36 
0.015-0.05 

5 

Jorgensen  (1976) 
Sheng  et  al.  (2001) 

Table  4.9  Water  quality  parameters  in  the  nitrogen  dynamics 

Paramete 
r 

description 

Unit 

literature 

value 

Reference 

K-onm 

Ammonification  rate 

1/day 

0.001-0.4 

0.1 

0.015 

0.01 

Bowie  et  al.  (1985) 

Pribble  et  al.  (1997) 

Cerco  and  Thomas 

(1995) 

Sheng  etal.  (2001) 

73 


KNN 

nitrification  rate 

1/day 

0.004-0.11 

0.08 

0.07 
0.01-0.02 

Bowie  et  al.  (1985) 
Pribble  et  al.(1997) 
Cerco  and  Thomas 
(1995) 
Shengetal.  (2001) 

Hnil 

DO  saturation  rate  for 
nitrification 

mg/1 

0.1-2.0 

2 
2 

Ambrose  (1994) 
Pribble  et  al.(1997) 
Sheng  et  al.  (2001) 

KDN 

denitrification  rate 

1/day 

0.02-  1.0 
0.09 
0.09 

Bowie  et  al.  (1985) 
Pribble  et  al.(1997) 
Sheng  et  al.  (2001) 

Hnoi 

DO  saturation  rate  for 
denitrification 

mg/1 

0.-  1.0 
0.1 

Bowie  et  al.  (1985) 
Shengetal.  (2001) 

Pm 

partition  coefficient  of 
PON/SON 

- 

l.E-5 
1.E-6-9E-6 

Simon  (1989) 
Sheng  et  al.  (2001) 

Pan 

partition  coefficient  of 
PIN/NH4 

5.E-6-1.E- 

5 
3.E-5-4.E- 

3 

Simon  (1989) 
Sheng  et  al.  (2001) 

don 

sorption/desorption  rate 
for  SON/PON 

1/day 

0.02 

0.08 

0.01-0.02 

Bowie  etal.  (1985) 
Cerco  and  Thomas 
(1995) 
Shengetal.  (2001) 

dan 

sorption/desorption  rate 
for  PIN/NH4 

1/day 

0.01 

Shengetal.  (2001) 

^■PDN 

preference  partition 
coefficient  of  mortality 
for  SON  /  PON 

- 

0.5 
0.5 

Cerco  and  Thomas 

(1995) 

Sheng  etal.  (2001) 

Table  4.10  Water  quality  parameters  in 

the  phos] 

Dhorous  dynamics 

Paramete 
r 

description 

Unit 

literature 

value 

Reference 

"~OPM 

mineralization  rate 

1/day 

0.001  -  0.6 

2.27 

0.1 

0.1 

Bowie  et  al.  (1985) 

Pribble  et  al.  (1997) 

Cerco  and  Thomas 

(1995) 

Sheng  et  al.  (2001) 

Pop 

partition  coefficient  of 
POP/SOP 

- 

8.E-6-1.E- 

4 

Shengetal.  (2001) 

74 


% 

partition  coefficient  of 
PIP/SRP 

- 

1.E-6-6E- 
4 

Sheng  et  al.  (2001) 

<p 

sorption/desorption  rate 
for  SOP/POP 

1/day 

0.08 
0.01 

Cerco  and  Thomas 

(1995) 

Sheng  etal.  (2001) 

dap 

sorption/desorption  rate 
for  PIP/SRP 

1/day 

0.01 

Sheng  etal.  (2001) 

PpDP 

preference  partition 
coefficient  of  mortality 
for  SOP  /  POP 

- 

0.5 
0.5 

Cerco  and  Thomas 

(1995) 

Sheng  etal.  (2001) 

Table  4.1 1  Water  quality  parameters  in 

the  oxygen  balance 

Paramete 
r 

description 

Unit 

literature 
value 

Reference 

KD 

oxidation  rate 

1/day 

0.02  -  0.6 
0.05 

Bowie  et  al.  (1985) 
Sheng  et  al.  (2001) 

"  CBOD 

DO  half  saturation  rate 
for  oxidation 

mg/1 

1.5-400 
0.5 

1 
10 

Bowie  et  al.  (1985) 

Pribbleetal.(1997) 

Cerco  and  Thomas 

(1995) 

Sheng  et  al.  (2001) 

J®CBOD 

partition  coefficient  of 

particular/dissolved 

CBOD 

- 

0.5 
0.3-0.5 

Bowie  et  al.  (1985) 
Sheng  et  al.  (2001) 

SOD 

Sediment  oxygen 
demand 

g02/m 

2-day 

0.02-10. 
0.0  -  10.7 

Thomann  (1972) 
Bowie  et  al.  (1985) 

"  SOD 

DO  half  saturation  rate 
for  sediment  oxygen 
demand 

mg/1 

0.01-0.22 
0.03-  0.09 
0.02-  0.06 

Jorgenson  (1979) 

Cerco  and  Thomas 

(1995) 

Sheng  et  al.  (2001) 

KAE 

reaeration  rate 

1/day 

0.  -  3. 

0.  -  0.25 

0.05-0.1 

Bowie  etal.  (1985) 

Cerco  and  Thomas 

(1995) 

Sheng  etal.  (2001) 

75 
Model  calibration  is  the  first  stage  testing  or  tuning  of  the  model  to  a  field  data  not 
used  in  the  original  construction  of  the  model.  Such  tuning  is  to  include  consistent  and 
rational  set  of  theoretically  defensible  parameters  and  inputs  (Thomann,  1992).  Proper 
calibration  of  the  water  quality  model  requires  having  accurate  representation  of  the  inflow 
and  loads  of  nutrients  into  the  water  body  and  selecting  appropriate  model  parameters. 

The  reaction  equation  shown  in  previous  section  is  a  function  of  its  concentration  and 
the  water  quality  parameters  connected  to  it  by  the  indicated  processes.  Within  each  reaction 
equation,  there  are  numerous  kinetic  parameters  and  additional  parameters.  Water  quality 
model  in  CH3D-EVIS  consists  of  13  state  variable  equations  with  over  40  interrelated 
parameters.  The  interactions  of  water  quality  model  parameters  and  the  constituent  equations 
as  shown  in  Table  4.12  clearly  indicate  the  complexity  of  the  calibrating  these  types  of 
models.  The  column  of  Table  4.12  show  that  each  modeled  constituent  equation  contains 
between  2  and  15  different  water  quality  parameters.  According  to  each  row,  each  parameter 
can  be  found  in  up  to  10  different  constituent  equations.  Therefore,  changing  one  parameter 
to  improve  the  calibration  of  one  constituent  will  simultaneously  affect  many  other 
constituents.  Traditionally,  calibration  of  water  quality  models  has  been  performed  manually 
using  a  trial-and-error  parameter  adjustment  procedure.  The  process  of  manual  calibration 
depending  on  the  number  of  model  parameters  and  the  degree  of  parameter  interaction.  With 
complexity  of  water  quality  model,  the  traditional  process  is  a  very  tedious  and  time 
consuming  task.  It  is  necessary  to  apply  more  systematic  and  efficient  calibration  procedure 
for  reducing  calibration  time  and  effort. 

Based  on  the  cascading  effect  of  adjusting  interrelated  parameters,  the  efficient 
calibration  of  the  water  quality  model  should  begin  with  the  parameters  that  affect  the  more 


76 
constituents  and  the  more  sensitive  parameters.  First  of  all,  each  water  quality  parameter 

can  be  ranked  with  these  sensitivity  and  relativity  as  parameterization.  According  to  this 

order,  high  ranked  parameters  will  be  adjusted  to  reproduce  major  pattern  of  all  constituents, 

and  then  lower  ranked  parameters  will  be  calibrated  for  detail  characteristics  of  local 

constituents.  This  procedure  will  reduce  numerous  calibration  efforts  and  increase  efficiency 

and  effectiveness. 

The  calibration  procedure  involves  optimization  of  numerical  measures  (objective 

functions)  that  compare  observations  of  the  state  of  the  system  with  corresponding  simulated 

values.  The  most  commonly  used  objective  function  adopted  in  calibration  is  the  root  mean 

square  errors  between  the  observed  and  simulated  model  response.  The  root  mean  square 

error  is  an  indication  of  the  average  discrepancy  between  observations  and  model  results. 

It  is  computed  as  follow: 


T(o-p)2 

RMS  =  J^- (4.45) 

n 
where  RMS  is  root  mean  square  error 

O  is  observation 

P  is  model  prediction 

n  is  number  of  observation 

In  addition  with  root  mean  square  error,  model  calibration  was  assessed  via  plots  of 
model  output  and  observation  with  correlation  coefficient  R2.  Scatter  plots  of  model  output 
and  observed  data  provide  an  indication  of  overall  model  performance.  The  correlation 
coefficient  is  defined  by 


77 


ssl  C£0P-nOP): 


R*  = 


(4.46) 


SS^-SS^     (£02-n.02).(£P2-n-P2) 
SSxx=Z(Ol-d)2=YJ02-n.02 
SS^ZiP.-Pf-TP'-n-P2 
SSv^iOt-OM-P^J^O-P-nO-P 

The  process  of  model  calibration  is  illustrated  in  Figure  4.7.  In  systematic  calibration 
procedure,  parameters  are  adjusted  according  to  order  of  sensitivity  and  relativity  according 
to  parameterization  for  optimization  of  certain  criteria  (objective  functions)  that  measure  the 
goodness-of-fit  of  the  simulation  model.  The  process  is  repeated  until  a  specified  stopping 
criterion  is  satisfied. 

Formation  of  a  proper  framework  for  systematic  calibration  involves  the  following 
key  elements: 

•  Sensitivity  analysis 

•  Model  parameterization  and  choice  of  calibration  parameters 

•  Specification  of  calibration  criteria 

The  best  way  to  calibrate  water  quality  parameters  is  the  automatic  calibration,  in 
which  parameters  are  adjusted  automatically  according  to  a  specific  search  scheme  for 
optimization  of  certain  calibration  criteria.  The  process  is  repeated  until  a  specified  stopping 
criterion  is  satisfied.  In  this  study,  automatic  calibration  procedures  have  been  developed 
and  tested,  using  a  Gauss-Newton  method. 

To  test  the  automatic  calibration  procedure  of  the  CH3D  water  quality  model,  a  1996 
baseline  run  was  adopted  as  representative  of  the  true  and  valid  field  condition.  All 
parameters  from  baseline  run  were  considered  representative  of  conditions  that  could 
hypothetically  exist  in  the  field.  A  small  number  of  these  parameters  were  perturbed  slightly, 
and  the  automatic  calibration  procedure  was  employed.   If  the  procedure  is  correct,  these 


78 
perturbed  parameters  should  converge  on  pre-perturbed  values.  The  results  of  the  test  show 
that  the  changed  parameters  did  converge  on  pre-perturbed  baseline  values,  with  the  Gauss- 
Newton  method.  The  method  is  therefore  considered  valid  and  accurate. 

In  real  case,  a  trial  run  was  conducted  by  using  all  the  measured  data  to  find  best  fit 
parameters.  The  calibration  procedure  continuously  failed  as  calibration  parameters  were 
automatically  adjusted  outside  the  upper  and  lower  bounds.  The  accuracy  of  the  model 
calibration  generated  by  this  procedure,  relies  heavily  on  the  quality  and  quantity  of  field 
observation,  the  model  structure,  and  the  nature  of  the  system.  Limited  data,  and 
uncertainties  in  water  quality  processes  caused  the  procedure  to  fail  to  generated  a  valid 
calibration  condition.  To  apply  an  automatic  calibration  procedure  to  this,  or  other  water 
quality  models,  further  investigation  is  required.  These  investigations  should  be  focus  on  the 
optimization  algorithm  and  refinement  of  the  water  quality  processes. 


79 


Table  4.12  The  relationship 

>  between  water  quality  parameters  and  model  constituents 

CA      CZ     PIN    NH4    N03    SON   PON    PIP     SRP    SOP    POP     DO    CBOD 

#of 
Eqns 

AGRM 
HALFN 
HALFP 
KAEX 
KAS 
WAS 
HALFA 
TRESHA 
ZGRM 
KZEX 
KZS 
SONM 
NITR 
DENR 
PCON 
DRON 
KPDN 
PCAN 
DRAN 
SOPM 
PCOP 
DROP 
KPDP 
PCIP 
DRIP 
SODM 
FD5 
AKD 
AOC 
CHLAC 
CAN 
ACP 
TOPT 
OPTL 
AKNIT 
AKDEN 
AKBOD 
AKSOD 
AKAIR 

1    i         ; 

!    l 

1      I 

l                     i    1    ; 

5 
1 
1 

5 

7 

1 

2 

2 

2 

3 

6 

2 

3 

2 

2 

2 

2 

2 

2 

2 

2 

2 

2 

2 

2 

1 

2 

2 

10 

2 

6 

5 

1 

1 

3 

2 

2 

1 

1 

i    i 

1"! i 

i    i                   l 

l    ! 

l 

l         l 

l    ; 

i   ■    '    : 

;    l    !     l 

l 

: 

:            : 

:              i              :              : 

i    I   i    i 

i        i    ; 

1    ;    i    ! 

! 

:              i              i              : 

i               l 

i 

i    i    i 

...........J 

l    i 

1       : 

1 

L.1...L.            ...L 

l 

l 

i   i 
i 

l 
l 

i 

1    i    1 

! 

1        1 

1   i 

1    i 

1 

i    1 

1 

l    i 

1             1 

1 

l 

: 

1 

l 

1 

l 

: 

l 

1 

i 

l 

1 

: 

"T'"? 

1 

1 

l 

i 

l 

l 

1 
j-    f    j-    1 

l        1    ! 
l    i    1    1 

1           1      ! 

i             11      ii 

11      =     1 

i 

111 



1    j 
l    j 

•1        i    ; 
!    l    ; 

t  "T"  "I 

t t m 

i       i 

1 ] 

!        !        I        !        !        i        i        i        !    i'T" 

#of 
Parameters 

15        82         10        7872887         14         7 

103 

80 


INITIAL  VALUE  for  WQ  parameter 


I 


Execute  WQ  model 


I 


Adjust  WQ  river  boundary  Condition 


I 


Execute  WQ  model 


change  i  parameter  for  sensitivity  test 


I 


Execute  WQ  model 


Sensitivity  analysis : 
check  RMS  differneces  between  base  run  and  each  sensitivity  test 


I 


parameterization  : 
make  order  with  sensitivity  result  and  relativity  table 


1 


Adjust  i  ranked  parameter 


Execute  WQ  model 


Calculate  statistics  :  scatter  plot,  RMS,  R2 
Plot  time  series  for  each  species  with  measured  datat 


Figure  4.7  Systematic  calibration  procedure 


CHAPTER  5 
APPLICATION  OF  CIRCULATION  AND  TRANSPORT  MODEL 

Circulation  in  Charlotte  Harbor  is  driven  by  tide,  wind,  and  density  gradient.  The 
numerical  model  of  circulation  and  transport  described  in  Chapter  3  was  applied  to  the 
Charlotte  Harbor  estuarine  system,  a  large  shallow  estuary  in  southwest  Florida.  Model 
applications  include  1)  the  simulation  of  three-dimensional  circulation  during  summer  1986 
2)one  year  simulation  of  flow,  salinity,  temperature,  and  sediment  transport  in  2000  and  3) 
simulation  of  the  impact  of  removal  of  Sanibel  Causeway  on  the  circulation  in  San  Carlos 
Bay  and  Pine  Island  Sound 

A  major  purpose  of  the  first  two  simulations  is  the  calibration  and  validation  of  the 
Charlotte  Harbor  circulation  and  transport  model  using  field  data  obtained  by  USGS,  NOAA, 
SWFWMD,  and  SFWMD  in  Summer  1986  and  January  to  December  in  2000.  During  the 
calibration  process,  a  few  model  coefficients  and/or  boundary/initial  conditions  are  adjusted 
to  produce  more  accurate  overall  simulation  of  observed  data  collected  in  July  1986  by 
USGS  (Hammet,  1992;  Stoker,  1992;  Goodwin,  1996)  covered  the  entire  Charlotte  Harbor 
estuarine  system  and  include  water  level,  horizontal  currents  and  salinity.  Hence  July  1986 
data  were  used  for  short-term  calibration.  To  supplement  the  short-term  calibration  of  July 
1986,  the  long-term  model  calibration  was  conducted  using  one  year  water  level  and  salinity 
data  in  Caloosahatchee  River  during  2000. 


81 


82 
5.1  A  High-Resolution  Curvilinear  Grid  for  Charlotte  Harbor  Estuarine  System 

To  procedure  a  successful  numerical  simulation  of  the  Charlotte  Harbor  estuarine 
system,  it  is  necessary  to  design  a  numerical  grid  which  represents  the  dominant  geographic/ 
bathymetric  features  with  sufficient  spatial  resolution.  Since  CH3D  use  a  boundary-fitted 
curvilinear  grid  in  the  horizontal  directions,  it  is  possible  to  align  the  grid  lines  to  coincide 
with  shorelines,  causeways,  and  bridges.  The  Charlotte  Harbor  estuarine  system 
encompasses  about  735  km2  on  Florida's  southwest  coast.  The  model  domain  includes  all 
the  sub-basins  (Upper  Charlotte  Harbor,  Lower  Charlotte  Harbor,  Pine  Island  Sound, 
Matlatcha  Pass,  and  San  Carlos  Bay),  the  major  tributaries  (Peace,  Myakka,  and 
Caloosahatchee  up  to  the  Franklin  Dam),  Estero  Bay,  and  some  offshore  water.  The  model 
should  be  run  with  a  spatial  grid  sufficiently  fine  to  resolve  the  Sanibel  Causeway.  The 
eastern  boundary  includes  many  of  the  major  rivers  with  specified  flow  rates,  while  the 
western  boundaries  and  the  western  portion  of  north  and  south  boundaries  are  open  tidal 
boundaries  where  tidal  elevations  are  prescribed.  The  interior  of  the  model  domain  is 
represented  with  a  boundary-fitted  grid  which  is  non-orthogonal  but  as  orthogonal  as 
possible,  i.e.  the  grid  analysis  are  usually  between  60°  and  120°. 

The  final  boundary  fitted  grid  used  for  the  three-dimensional  curvilinear-grid  model 
(CH3D)  of  Charlotte  Harbor  estuarine  system  was  generated  using  a  grid  generation  program 
originally  developed  by  Thompson  et  al.  (1985)  and  further  enhanced  for  this  study.  This 
grid  (Figure  5.1)  contains  92  x  129  horizontal  cells  and  eight  vertical  layers,  with  a  total  of 
1 1 648  horizontal  grid  cells  which  include  5367  water  cells  and  628 1  land  cells.  Grid  spacing 
varies  from  40  to  2876  meters  (average  598  meters). 

The  bathymetry  of  the  Charlotte  Harbor  estuarine  system  and  the  nearshore  region 


83 
of  the  Gulf  of  Mexico  is  derived  from  data  obtained  from  the  Geophysical  Data  System  of 
National  Geophysical  Data  Center.  Bathymetry  of  navigation  channels  in  San  Carlos  Bay 
and  the  vicinity  of  Sanibel  Causeway  are  based  on  the  recent  survey  data  provided  by  Lee 
County  in  December  1 999.  Bathymetry  for  Peace  River  and  the  upper  Charlotte  Harbor  area 
was  collected  by  SWFWMD  and  that  for  Caloosahatchee  River  was  collected  by  SFWMD. 
All  bathymetric  data  were  converted  to  NAVD88  to  unify  vertical  datum  level  (Appendix 
H).  Charlotte  Harbor  grid  bathymetry  (Figure  5.2)  was  developed  using  all  these  bathymetric 
data.  While  an  inverse  distance  interpolation  followed  by  simple  smoothing  scheme  was  the 
primary  method  for  determining  bathymetry  in  the  study  area,  several  areas  were  added  and 
the  bathymetry  further  adjusted  after  the  interpolation  and  smoothing  was  performed  to 
ensure  the  proper  passage  of  flow  through  a  navigation  channels. 


84 


C 

o 

z 


\ 

J? 

? 

A     V 

o 

h 

^^fc^ffiE^ 

o 

Sv»T- 

o 

AVC*^ 

in 

\yvow**' 

Y     ,--- 

r- 

co 

v" 

CM 

j  i^^H 

O 

o 

o 

o 

^m                     \s&& 

LO 

CD 

SgS&SRi  s^^W 

^■rstLrf/fiv 

C\J 

w 

§11  \1 

o 

illlllll 

WELug;        \  L 

o 

?Crt>,   Vv^H| 

o 

■~''s$5r>SJtit.  4$K2> 

LO 

— 

C\J 

en 

SSsc$v---" 

CM 

v'     '' '~^*^$x?xX& 

il 

o$x38BKv\S^^^^S^Stfcjj*t--> 

v^^sSSSKS^^^^ww^ 

o 

o 

o 

s/?c^>otjo?v^0cfilS^^B^ 

o 
o 

CO 

liPH 

CM 

lilt 

i     i 

1 

1           1       _J_S^ 

r     1      i      i      i      i      1      i      i      i      i 

350000 


375000  400000 

Easting  (m) 


425000 


Figure  5.1  Boundary-fitted  grid  (92  x  129)  used  for  numerical  simulation  for  Charlotte 
Harbor  estuarine  system. 


85 


£      H>£J 


r 


J I I L 


J i i i i 


350000 


375000  400000 

Easting  (m) 


425000 


2000 

1900 

1800 

1700 

1600 

1500 

1400 

1300 

1200 

1100 

1000 

900 

800 

700 

600 

500 

400 

300 

200 

100 

0 


Figure  5.2  Bathymetry  in  boundary-fitted  grid  for  Charlotte  Harbor  estuarine  system  (92 
x  129). 


86 
5.2  Forcing  Mechanism  and  Boundary  Conditions 

The  forcing  mechanisms  for  the  hydrodynamic  model  include  tide,  wind,  and  density 
gradients.  To  incorporate  these  mechanisms  in  the  model  simulations,  tides  and  oceanic 
salinity  are  specified  along  the  offshore  boundary,  wind  is  specified  at  the  air-sea  interface, 
and  river  discharges  are  specified  at  the  river  boundaries. 

The  boundary  between  the  Charlotte  Harbor  estuarine  system  and  the  Gulf  of  Mexico 
extends  about  64  km  from  Gasparilla  Pass  on  the  north  to  San  Carlos  Pass  on  the  south.  One 
of  the  most  important  features  is  the  similarity  between  the  tidal  curves  at  Venice  and  Cayo 
Costa  which  are  located  at  northern  end  and  near  Boca  Grande  Pass  (Figure  1.1).  The 
phasing  of  the  Gulf  tides  is  nearly  synchronous  along  the  boundary  of  the  estuarine  system 
between  Venice  and  Naples,  which  are  located  at  the  northern  and  southern  ends. 

For  the  2000  simulation,  tidal  forcing  along  the  Gulf  of  Mexico  boundary  was 
prescribed  with  water  level  data  measured  at  the  Naples  station  by  the  NOAA  CO-OPS.  For 
the  1986  simulation,  water  level  data  at  Venice  measured  by  USGS  were  used  to  provide  the 
tidal  boundary  condition,  since  there  was  no  data  from  the  Naples  station  in  this  period.  To 
unify  the  datum  level  with  the  bathymetry,  these  water  level  data  were  converted  to  NAVD88 
datum  level. 

Discharge  and  runoff  boundary  conditions  were  imposed  using  the  daily  measured 
discharge  data  described  in  Table  4.1.  Because  no  flow  data  were  available  for  the  Estero 
Bay  and  Caloosahatchee  River  except  at  S-79  in  1986,  the  discharges  were  assumed  to  be 
one  half  and  a  quarter  of  the  Myakka  River  discharges,  respectively. 

The  surface  wind  boundary  condition  is  supplied  using  hourly  wind  magnitude  and 
direction  data  collected  at  three  stations,  which  are  the  National  Data  Buoy  Center  C-MAN 


87 
stations  at  Venice  (27.07  N,  82.45  W  ),  NOAA  CO-OPS  stations  at  Fort  Myers  (26°  36'  N, 
81°  52'  W),  and  Naples  (26.13  N,  81.807  W  ).  For  the  2000  simulation,  hourly  wind  data 
from  all  three  stations  are  available,  while  only  the  Venice  and  Fort  Myers  stations  are 
available  for  1986  simulation.  The  hourly  wind  magnitude  and  direction  are  converted  into 
x-(EastAVest)  and  y-  (North/South)  velocity  components  using  Garrat's  formula  (Garrat, 
1977).  These  values  are  then  interpolated  to  the  entire  computational  grid,  with  a  weighting 
function  inversely  proportional  to  the  square  of  distances.  Figures  5.3  shows  the  water  level 
at  Venice  station  for  tidal  boundary  condition  and  the  river  flows  at  river  boundaries  in  1986. 
Figure  5.4  shows  wind  speed  and  direction  at  Venice  and  Fort  Myers  station  in  1986.  Figure 
5.5  shows  the  water  level  at  Naples  station  for  tidal  boundaries  and  the  river  flows  at  river 
boundaries  in  2000.  Figure  5.6  shows  wind  speed  and  direction  at  Venice,  Fort  Myers  and 

Naples  in  2000. 

Table  5.1  Descriptions  of  1986  and  2000  river  boundary  conditions  for  Charlotte  Harbor 


wluu""v  "J" 
Station 

Name 

Agency 

Latitude 

Longitude 

drainage 

perio 

degree  W 

degree  N 

area:mi2 

d 

Myakka  River 

02298830 

near  Sarasota 

USGS 

27°14'25" 

82°18'50" 

229.0 

daily 

02299120 

at  Deer  Prairie  Slough 

USGS 

27°08'06" 

82°15'24" 

- 

daily 

02299410 

at  Big  Slough  Canal 

USGS 

27°11'35" 

82°08'40" 

36.5 

daily 

Peace  River 

02296750 

at  Arcade 

USGS 

27°13'19" 

81  °52'34" 

1 ,376.0 

daily 

02297100 

at  Joshua  Creek 

USGS 

27°09'59" 

81°52'47" 

132.0 

daily 

02297310 

at  Horse  Creek 

USGS 

27°11'57" 

81°59'19" 

218.0 

daily 

02298202 

Shell  Creek 

USGS 

26°59'04" 

81  °56'09" 

373.0 

daily 

in  Caloosahatchee  River 

02292900 

S79  spillway 

USGS 

26°43'25" 

81°41'55" 

- 

daily 

02293214 

Cape  Coral 

USGS 

26°38'10" 

81  °5548" 

- 

daily 

02293230 

Whiskey  Creek 

USGS 

26°34'29" 

81  °53'29" 

" 

daily 

in  Estero  Bay 

02291500 

Imperial 

USGS 

26°20'07" 

81°44'59" 

65.0 

daily 

02291524 

Spring  Creek 

USGS 

26°21'42" 

81  °47'27" 

- 

daily 

02291580 

Estero  River 

USGS 

26°26'30" 

81°47'45" 

- 

daily 

02291673 

Mullock  &  Henry  Creek 

USGS 

26°30'19" 

81°51'00" 

~ 

daily 

88 
Baroclinic  simulations  require  realistic  initial  salinity  and  temperature  fields  in  order 
to  minimize  the  effects  of  the  initial  conditions  on  the  solution,  and  to  minimize  the  "spin- 
up"  time.  Salinity  along  the  offshore  boundary  is  assumed  to  be  approximately  constant  at 
36.5  ppt.  Salinity  at  the  river  boundaries  was  set  to  0  ppt.  The  near  bottom  salinity  data 
measured  at  the  nine  USGS  stations  in  1986  were  used  to  determine  the  initial  salinity  at  the 
corresponding  grid  cells.  These  salinity  data  were  interpolated  over  the  entire  computational 
grid,  with  a  weighting  function  inversely  proportional  to  the  square  of  the  distance  to  the 
three  nearest  stations.  Initial  temperature  field  for  2000  simulation  was  interpolated  using 
water  temperature  data  at  Fort  Myers  and  Naples  stations  measured  by  NOAA  CO-OPS  and 
Venice  stations  measured  by  National  Data  Buoy  Center  (NDBC).  Air  temperature  data  at 
the  same  locations  were  used  to  calculate  heat  flux  for  air-sea  interface  (Figure  5.7).  These 
air  temperature  data  were  also  interpolated  over  the  study  area. 


89 


160 


1 80  200 

Julian  Day 


220 


300  r 


Caloosahatchee  River 
Peace  River 
Shell  Creek 
Myakka  River 


160 


180  200 

Julian  Day 


220 


Figure  5.3  Tidal  forcing  and  river  discharges  for  1986  simulation  of  Charlotte  Harbor 
circulation. 


at  Venice 


90 


I     ; 


'     I     I 


I     i 


150  160  170  180  190  200 

Julian  Day 


210 


220 


230 


at  Fort  Myers 


j_ 


150  160  170  180  190  200 

Julian  Day 


210 


220 


230 


Figure  5.4  Wind  velocity  for  1986  simulation  of  Charlotte  Harbor  circulation. 


91 


Julian  Day 


200  r 


100 


Caloosahatchee  River 
Peace  River 
Shell  Creek 
Myakka  River 


200 

Julian  Day 


300 


Figure  5.5  Tidal  forcing  and  river  discharges  for  2000  simulation  of  Charlotte  Harbor 
circulation 


Wind  speed  and  direction  at  Venice 


92 


36850 


36875 


36900 


36925 


36950 


Wind  speed  and  direction  at  Fort  Myers 


36850 


36875 


36900 


36925 


36950 


Wind  speed  and  direction  at  Naples 


36660 


36720 


36680  36700 

Julian  Day  since  1 900 

Figure  5.6  Wind  velocity  for  2000  simulation  of  Charlotte  Harbor  circulation. 


93 


at  Venice 


100 


200 

Julian  Day 


300 


at  Fort  Myers 


200 

Julian  Day 


100  200  300 

Julian  Day 

Figure  5.7  Air  temperature  for  2000  simulation  of  Charlotte  Harbor  circulation. 


94 
5.3  Simulations  for  1986  Hydrodynamics 

During  the  calibration  process,  a  few  model  coefficients  and/or  boundary  and  initial 
conditions  were  adjusted  to  allow  accurate  simulation  of  observed  data.  As  shown  in  Table 
5.2,  measured  data  of  tidal  stage,  current,  and  salinity  at  several  locations  throughout  the 
Charlotte  Harbor  were  provided  by  USGS  (Goodwin,  1992).  Figure  5.8  shows  the  locations 
of  tidal  stages  and  Figure  5.9  shows  the  locations  of  current  velocity  and  salinity 
measurement.  Water  elevation  data  at  USGS  station  4  (Venice)  were  compared  with  those 
at  USGS  stations  5  (Cayo  Costa)  and  8  (Gulf  of  Mexico  at  Fort  Myers)  which  are  located  in 
the  Gulf  of  Mexico  (Figure  1.1).  The  tidal  elevation  and  phase  at  stations  4  and  5  are  almost 
identical.  Tidal  range  at  Fort  Myers  beach  (station  8)  is  larger  than  those  at  stations.  The 
reason  for  the  larger  tidal  range  at  Fort  Myers  is  probably  due  to  the  abrupt  change  in  the 
direction  of  shoreline  at  the  southern  end  of  Sanibel  Island,  rather  than  different  offshore 
tides.  Therefore,  water  level  data  at  Venice  station  were  used  as  tidal  boundary  conditions 
along  all  open  boundaries.  All  specific  conductance  measurements  were  converted  to  salinity 
concentrations  in  parts  per  thousand  using  the  conversion  equation  by  Miller  (Goodwin. 
1992).  These  data  were  used  for  comparison  with  model  simulations  of  water  level,  velocity 
and  salinity. 

Before  baroclinic  simulations  are  performed,  it  is  necessary  to  generate  an  initial 
salinity  field  because,  in  general,  salinity  is  much  slower  to  adjust  to  initial  transients  than 
water  level  or  currents.  The  initial  salinity  field  for  the  1986  simulation  is  generated  by 
"spinning  up"  a  prescribed  salinity  field  for  a  sufficiently  long  period.  The  salinity  values 
at  the  end  of  the  simulation  are  then  used  the  initial  values  for  the  1986  simulations.  This 
process  is  discussed  below  in  more  detail. 


95 

First,  a  salinity  field  is  created  by  linearly  interpolating  measured  salinity  onto  the 
Charlotte  Harbor  grid.  The  bottom  salinity  data  measured  by  USGS  during  1986  are  used 
to  create  this  salinity  field.  Next,  a  salinity  assimilation  term  is  added  to  the  salinity  transport 
equation  in  the  model.  This  term  forces  the  salinity  in  a  selected  cell  to  approach  a  specific 
value,  in  this  case,  the  initial  value.  The  six  USGS  measured  bottom  salinity  values  shown 
in  Figure  5.6  are  chosen  to  be  assimilated  into  the  simulation.  These  values  were  chosen 
because  they  were  measured  closest  to  the  July  9  starting  date  of  the  1986  simulations.  The 
assimilation  term  ,  which  appears  on  the  right  side  of  the  salinity  equation,  takes  on  the 
following  form  (Sheng  and  David,  2002): 

T 

where  S"iJk  is  the  simulated  salinity  at  the  n-th  time  level,  (SA)iJk  is  the  value  to  be 

assimilated,  i.e.,  the  initial  salinity  values,  and  T  is  the  assimilation  period,  which  is  set  to 
30  days  for  Charlotte  Harbor  simulation.  It  can  be  seen,  when  S"ijk  is  not  equal  to  (SA)iJk  a 
force  is  created  to  drive  S"iJk,  toward  the  value  of  (SA)ijk.  Because  of  the  addition  of  the 
salinity  assimilation  term,  the  salinity  at  the  six  USGS  sites  at  the  end  of  a  simulation  will 
be  very  close  to  their  measured  values  at  the  beginning  of  the  simulation. 

A  30-day  spin-up  simulation  was  performed  from  May  28  to  June  27  during  a  dry 
season  with  all  barotrophic  forcing  mechanisms  (tides,  river  discharges,  wind)  to  allow  water 
level,  velocity  and  salinity  field  to  reach  a  dynamic  steady-state  throughout  the  computational 
domain.  Using  the  last  time  step  surface  elevation,  velocity,  and  salinity  of  the  spin-up 
simulation  as  the  initial  condition,  the  Charlotte  Harbor  circulation  and  salinity  transport 
from  June  27  to  July  30  was  simulated  with  tidal  forcing,  wind  field  and  river  discharges  as 
boundary  condition. 


96 


Site 

(I  ,J)  Location  in 

Duration 

Data  Type 

Number 

Computational 
domain 

1 

(29,106) 

6/15-8/31,1986 

Tidal  Stage 

2 

(40,102) 

7/15-8/31,1986 

Tidal  Stage 

3 

(23,  81) 

6/15-8/31,  1986 

Tidal  Stage 

4 

Outside,  at  Venice 

6/30-8/31,1986 

Tidal  Stage 

5 

(12,  81) 

6/30-8/31,  1986 

Tidal  Stage 

6 

(14,  62) 

7/10-8/31,1986 

Tidal  Stage 

7 

(28,  47) 

8/14-8/31,1986 

Tidal  Stage 

8 

(33,  29) 

6/26  -  8/20,  1986 

Tidal  Stage 

SI-1 

(33,100) 

7/9-8/  6,  1986 

Velocity  &  Salinity 

SI-2 

(30,  90) 

11  9-8/  8,  1986 

Velocity  &  Salinity 

SI-3 

(35,  86) 

7/9-8/  6,  1986 

Velocity  &  Salinity 

SI-4 

(11,  96) 

7/10-7/15,  1986 

Velocity  &  Salinity 

SI-6 

(18,  76) 

7/10  -  7/23,  1986 

Velocity  &  Salinity 

SI-7 

(17,  65) 

7/10  -  7/23,  1986 

Velocity  &  Salinity 

SI-8 

(20,  51) 

7/16  -  7/20,  1986 

Velocity  &  Salinity 

97 


o 
o 
o 
o  I— 

o 

o 

CO 


o 

o 
o 
m 

(j> 

CM 


£  8 

w  o 

£    CM 

o 


o 
o 
o 
o 
o 
a> 

CM 


_ 


J_ 


350000 


•  Tidal  Measurement 
Stations  (1  to  8) 

0   Discharge  Measurement 
Stations  (9  to  15) 


ite 


r 


4 


100 


Y>Ur 


7  ^# 


J I . L 


J_ 


J L 


J I I L 


J__l L 


J_ 


375000 


400000 

Easting  (m) 


425000 


450000 


Figure  5.8  Locations  of  1986  water  level  and  discharge  measurement  stations  of  USGS 
(Goodwin,  1992). 


98 


o 
o 
o 
o 
o 
o 

CO 


w  o 

U)  o 

.b  en 

.C    CM 

t 

o 

z 

o 
o 
o 
to 

CM 

en 

CM 


o 
o 
o 
o 
o 
en 
c\j 


J L 


± 


Velocity  and  Salinity 
Measurement  Stations 
(SI-1  toSI-10) 


SI-3    / 


s|-9\  %¥%&* 


V 


SI-10 


J I l_L 


J_ 


J I L 


_L 


J I I L 


-L 


J__l I L 


350000      375000      400000      425000      450000 

Easting  (m) 


Figure  5.9  Locations  of  1986  velocity  and  salinity  measurement  stations  of  USGS 
(Goodwin,  1992). 


99 
5.3.1  Sensitivity  and  Calibration  Simulations 

Several  simulations  were  performed  in  order  to  test  the  consistency  and  sensitivity 
of  numerical  grid,  bottom  roughness,  horizontal  diffusion  coefficient,  and  bathymetry.  To 
quantify  the  model  sensitivity  to  variations  of  these  model  parameters,  time  series  statistics 
are  presented  in  terms  of  the  root-mean-square  error  (RMS)  and  normalize  RMS  error,  define 
as  the  ratio  between  the  RMS  error  and  the  observed  range.  The  normalized  RMS  error  gives 
a  more  meaningful  indication  on  model's  ability  to  reproduce  the  tidal  signal  and  salinity  at 
each  station.  The  RMS  error  is  calculated  according  to: 


E = 


■    ,      N  -11/2 

,r   /  j  \     simulated  data  ) 

TV       i 


(5.2) 


where  N  is  the  total  number  of  measured  data,  Ssimulated  is  the  model  result,  and  Sdata  is 

measured  data  at  each  station. 

To  determine  which  non-tidal  boundary  and  forcing  conditions  are  most  important 

to  Charlotte  Harbor  estuarine  system  modeling,  a  series  of  1986  simulations  were  performed 

with  a  different  boundary  and  forcing  conditions  removed  in  each  run  (Table  5.3).  These 

simulations  show  that  baroclinc  forcing  is  the  most  important  factor  in  simulating  water 

level,  flow,  and  salinity  at  overall  available  measured  stations. 

Table  5.3  The  effect  of  removing  selected  boundary  conditions  on  the  accuracy  of  simulated 
water  level,  velocity  and  salinity  in  July  1986.  Values  shown  are  average  RMS  differences 
vs.  baseline  simulation  at  all  data  stations. 


Boundary  and  Forcing 

Water  level 

U  Velocity 

V  Velocity 

Salinity  (ppt) 

Condition  Removed 

(cm) 

(cnr/s) 

(cnr/s) 

Baroclinic  Forcing 

2.35 

3.17 

2.05 

3.08 

Wind 

1.82 

2.64 

2.08 

1.03 

River  Discharge 

0.57 

1.15 

0.84 

1.05 

As  shown  in  Equation  3.28,  the  bottom  drag  coefficient  is  defined  as  a  function  of 


100 
the  bottom  roughness,  z0.  The  default  value  of  bottom  roughness,  z0,  used  for  Charlotte 
Harbor  simulations  is  chosen  to  be  a  constant  0.3  cm  over  the  entire  domain  based  on 
running  with  the  several  bottom  roughness  values  which  range  between  0.01  and  1  (Table 
5.4).  Because  of  the  numerous  different  bottom  types  in  the  estuary,  a  constant  value  of 
bottom  roughness  may  not  be  appropriate.  To  determine  whether  a  spatially  varying  bottom 
roughness  or  a  constant  bottom  roughness  produces  better  simulated  circulation  and 
transport,  a  series  of  1998  simulation  was  performed  for  the  Indian  River  Lagoon  (Davis  and 
Sheng,  200 1 ).  The  results  show  that  simulations  with  the  spatially  varying  bottom  roughness 
have  slightly  smaller  errors  in  water  level  but  slight  larger  errors  in  salinity  and  flow.  Since 
the  varying  bottom  roughness  does  not  improve  the  simulated  circulation  and  transport 
significantly,  the  constant  bottom  roughness  of  0.3  cm  is  used  for  all  simulations  in  this 

study. 

Table  5.4  The  effects  of  varying  bottom  roughness,  z0,  on  the  accuracy  of  simulated  water 
level  velocity,  and  salinity  in  July  1986.  Values  shown  are  average  RMS  errors  at  all  data 
stations. 


Bottom  roughness  (cm) 

Water  level 
(cm) 

Velocity 
(cm:/s) 

Salinity  (ppt) 

0.1 

8.11 

7.87 

1.71 

0.3 

8.12 

7.88 

1.71 

1 

8.11 

7.87 

1.71 

As  shown  in  Equations  3.2  and  3.3,  the  sub-grid  scale  motion  is  estimated  with  a 
horizontal  diffusion  coefficient,  AH.  A  default  value  of  10,000  cm2/s  was  used  for  Charlotte 
Harbor  simulations  after  performing  several  simulations  using  different  coefficient  values 
(Table  5.5).  The  results  show  little  difference  in  water  level  and  salinity. 

Besides  adjusting  boundary  conditions  and  model  coefficients,  it  is  also  important  to 
use  more  accurate  grid  and  bathymetry.  To  improve  the  accuracy  of  bathymetry  in  Charlotte 


101 

Harbor,  bathymetric  surveys  were  performed  in  Caloosahatchee  River  and  San  Carlos  Bay, 

and  upper  Charlotte  Harbor  by  SFWMD  and  SWFWMD,  respectively. 

Table  5.5  The  effect  of  varying  horizontal  diffusion,  AH,  on  the  accuracy  of  simulated  water 
level,  velocity  and  salinity  in  July  1986.  Values  shown  are  average  RMS  errors  at  all  data 


Horizontal  diffusion 
Coefficient  (cm2/s) 

Water  level 

(cm) 

Velocity 
(cm2/s) 

Salinity  (ppt) 

5000 

8.12 

7.86 

1.75 

10000 

8.12 

7.88 

1.71 

20000 

8.07 

7.88 

1.64 

Overall,  the  model  is  able  to  simulate  the  surface  elevation  and  salinity  within  10% 
normalized  error  with  maximum  value.  This  provides  validation  that  the  hydrodynamic 
model  reproduces  the  basic  circulation  of  the  Charlotte  Harbor  estuarine  system. 

To  supplement  this  short  term  calibration  of  July  1986,  long  term  model  calibration 
was  conducted  using  one  year  water  level  and  salinity  data  in  Caloosahatchee  River  during 
2000.  This  2000  model  calibration  include  several  test  simulations  to  investigate  effects  of 
bottom  roughness,  salinity  advection  schemes,  and  grid  resolution  and  bathymetry. 
5.3.2  Results  of  the  1986  simulation 

For  the  1986  simulation,  the  92x129  grid  and  updated  bathymetry  were  used  along 
with  boundary  conditions  and  model  parameters  described  in  Table  5.6.  Water  level,  current 
velocity,  and  salinity  are  compared  both  qualitatively  and  quantitatively  with  measured  data, 
where  available. 

Water  Level 

Calculated  RMS  errors  between  simulated  and  measured  water  level  during  the  July 
1986  simulation  are  shown  in  Table  5.7.  The  normalized  RMS  errors  are  less  than  10%  at 
all  water  level  stations,  demonstrating  the  model's  ability  to  accurately  reproduce  surface 


102 

elevation  in  the  system.   The  highest  errors  at  stations  1  and  2  can  be  attributed  to  the 

relatively  coarse  horizontal  grid  near  the  river  where  the  gages  were  located. 

Table  5.6    A  summary  of  boundary  conditions  and  model  parameters  used  in  the  1986 

simulation. . 

Boundary  Condition  or  Model  Parameter  Value 

Tidal  Forcing  Measured 

Wind  speed  and  direction  Measured  at  2  stations 

Fresh  Water  Discharge  Estimated  (for  Estero  Bay)  and  Measured 

Bottom  Roughness  Constant  (0.3  cm) 

Horizontal  Diffusion  Constant  (10000  cm2/s) 

Horizontal  grid  92  x  129 

Vertical  Layers 8 


Table  5.7  Calculated  RMS  errors  between  simulated  and  measured  water  level  in  July  1986. 
Station  Number  RMS  error  (cm)  Range  (cm)  %  RMS  error 


1 

9.59 

104.24 

9.20 

2 

9.61 

104.55 

9.19 

3 

6.61 

90.83 

7.28 

5 

9.55 

117.05 

8.16 

6 

7.56 

107.90 

7.00 

7 

5.78 

144.48 

4.00 

Average 

8.12 

111.51 

7.47 

Figure  5.10  shows  the  comparison  between  measured  and  simulated  water  level  at 
six  Charlotte  Harbor  stations.  The  result  shows  a  good  agreement  between  data  and  model 
results,  in  both  amplitude  and  phase.  From  these  figures,  it  can  be  seen  that  tidal  range 
decreases  about  15%  from  Boca  Grande  Pass  (site  5)  to  Peace  and  Myakka  Rivers  (sites  1 
and  2).  Also,  the  tidal  wave  takes  about  two  to  three  hours  to  reach  these  rivers. 

Figure  5.11  shows  the  spectra  of  water  level  (data  and  model  results)  for  the  same 


103 
stations.  The  spectral  curves  reveal  major  energy  bands  centered  around  two  tidal 
frequencies:  the  diurnal  band  (between  0.8  to  1.2  cycle  per  day)  and  a  semi-diurnal  band  (1.8 
to  2.2  cycles  per  day).  These  two  bands  combined  to  represent  the  tides  propagating  from 
Gulf  of  Mexico  into  the  estuarine  system.  The  secondary  energy  bands,  representing 
nonlinear  interactions  between  diurnal  and  semi-diurnal  tides  with  complex  geometries  and 
bathymetric  features  inside  the  estuary,  are  represented  by  the  relatively  small  third  and 
fourth  diurnal  peaks.  Comparisons  show  that  the  model  is  able  to  capture  both  the 
amplitudes  and  phases  well. 

Velocity 

The  simulated  and  measured  horizontal  velocities  at  selected  sites  are  shown  in 
Figure  5.12.  The  angle  of  the  vector  shows  the  azimuth  of  the  flow  vector  at  a  specific  time, 
and  the  length  of  the  vector  represents  the  magnitude  of  the  flow.  It  is  important  to  point  out 
that  the  observed  velocity  is  measured  at  a  discrete  point  in  space  and  is  subjected  to  many 
local  influences  that  are  not  well  represented  in  the  model.  These  influences  include  physical 
features  such  as  small  channels,  depression  shoals,  and  mounds  that  are  not  resolvable  at  the 
model  scale,  but  often  influence  the  distribution  of  velocity  at  that  location.  Because  of  local 
influences,  discrete  measurement  points  are  not  consistently  reliable  indicators  of  the  general 
velocity  characteristics  (Goodwin,  1996).  The  model  results  are  spatially  averaged  values 
over  a  grid  cell,  hence  are  not  expected  to  agree  exactly  with  discrete  data. 

Site  SI-1  is  located  in  northern  Charlotte  Harbor  in  an  area  of  converging  river  flow 
and  having  a  complex  bathymetry  that  creates  a  complex  lateral  distribution  of  velocity.  The 
simulated  velocity  at  SI-1  agrees  well  with  measured  data.  However,  some  simulated 
velocity  magnitudes  were  50%  smaller  than  measured  magnitudes  and  there  was  some 


104 
deviation  in  the  azimuth  of  more  than  30  degrees.  Since  the  current  meter  was  located  near 
the  surface  at  this  station,  this  station  could  be  influenced  by  freshwater  inflow,  wind  and  any 
other  effects  as  indicated  by  Goodwin  (1996). 

At  Si-2  and  Si-3,  there  is  good  agreement  between  simulated  and  measured  velocities. 
In  the  Pine  Island  Sound,  site-7  is  located  in  the  middle  of  several  islands,  and  the  observed 
data  were  strongly  affected  by  local  features  that  cannot  be  resolved  at  the  scale  modeled. 
However,  agreement  between  simulated  and  measured  data  for  this  site  is  quite  good. 

Calculated  RMS  errors  between  simulated  and  measured  current  velocity  are  shown 
in  Table  5.8.  The  average  normalized  RMS  error  is  less  than  20%.  The  RMS  error  is  quite 
reasonable  in  comparison  to  the  range  of  velocity,  considering  the  fact  that  many  local 
influences  that  are  not  well  represented  in  the  model. 
Table  5.8  Calculated  RMS  errors  between  simulated  and  measured  current  velocity  for  1986 


Station 

RMS  error  (cm/s) 

Range  (cm/s) 

%  RMS  error 

Number 

U 

4.38 

14.34 

SI-1 

V 

5.97 

30.54 

19.56 

u 

11.70 

21.20 

SI-2 

V 

13.14 

55.19 

23.82 

V 

7.16 

18.72 

SI-3 

V 

6.78 

38.25 

17.72 

u 

3.06 

8.79 

SI-6 

V 

10.82 

34.79 

31.10 

Average 

7.88 

19.40 

Salinity 

Figure  5.13  shows  the  near-bottom  simulated  and  measured  salinity  at  USGS  stations 


105 

SI-1,2,3,6,7,  and  8  inside  of  Charlotte  Harbor.    The  difference  between  simulated  and 

measured  salinity  does  not  exceed  2  ppt  except  at  station  SI-7. 

Calculated  RMS  errors  between  simulated  and  measured  salinity  during  July  1986 

simulation  are  shown  in  Table  5.9.  The  results  show  the  model's  ability  to  simulate  salinity 

within  10%  accuracy.  The  highest  error  is  at  station  SI-7.  The  measured  salinity  shows  very 

strong  daily  fluctuation  which  is  five  times  greater  than  simulated  values.  This  location  is 

near  the  Captiva  pass  which  is  possibly  affected  by  local  features  such  as  small  channels, 

depressions,  shoals,  and  mounds  as  mentioned  before. 

Table  5.9  Calculated  RMS  errors  between  simulated  and  measured  salinity  for  1986 

simulation . 

Station  Number  RMS  error  (ppt)  Maximum  (ppt)  %  RMS  error 


SI-1 

1.23 

24.01 

5.15 

SI-2 

1.47 

25.42 

5.77 

SI-3 

1.51 

27.05 

5.60 

SI-6 

0.61 

33.83 

1.81 

SI-7 

3.52 

33.82 

10.42 

SI-8 

1.90 

25.38 

7.50 

Average 

1.71 

26.55 

6.04 

Flow  Patterns 

Typical  averaged  flow  patterns  during  one  tidal  cycle,  based  on  the  water  level  at 
open  boundary,  for  August,  6  on  1986  are  shown  in  figure  5.14.  Easterly  flow  on  the  western 
Florida  shelf  is  dominant  during  the  flood  tide,  whereas  northerly  and  westerly  shelf  flows 
are  dominant  during  the  ebb  tide.  Two  major  passes  in  the  Charlotte  Harbor  estuarine  system 
are  Boca  Grande  Pass  and  San  Carlos  Pass.  The  large  flood  and  ebb  flows  through  the  Boca 
Grande  Pass  affects  the  tidal  prism  in  upper  and  lower  Charlotte  Harbor  and  the  flow  through 


106 
San  Carlos  Pass  satisfies  the  tidal  prism  in  the  Caloosahatchee  River,  San  Carlos  Bay,  the 
lower  part  of  Matacha  Pass  and  lower  extremity  of  Pine  Island.  The  flow  through  Gasparilla, 
Captiva,  and  Redfish  passes  have  effects  that  appear  to  be  limited  to  the  local  area.  In  Pine 
Island  Sound,  the  flow  is  very  low  during  flood  and  ebb  tide  periods.  However,  there  is  some 
water  transport  through  Pine  Island  Sound  during  high  slack  and  low  slack  tide  periods. 
During  near  high  slack  tide,  there  is  northerly  flow  from  San  Carlos  Bay,  while  there  is 
southerly  flow  during  near  low  slack  tide. 

Residual  Flow  and  Salinity  Patterns 

Figure  5.15  show  the  29-day  residual  circulation  during  July  2  to  July  30,  1986.  In 
the  upper  Charlotte  Harbor,  there  are  opposite  direction  surface  residual  flows  which  are  the 
landward  surface  residual  flow  in  relatively  deep  channel  along  the  right  side  channel  and 
seaward  surface  residual  flow  along  the  left  side  shoreline.  In  the  bottom  residual  flow,  the 
landward  flow  is  dominant  across  the  channel. 

In  the  Pine  Island  Sound  and  Matlacha  Pass,  the  northward  residual  flow  is  dominant 
in  the  surface,  while  there  is  a  week  southward  flow  in  the  bottom.  The  surface  flow  come 
from  upper  Charlotte  Harbor,  Pine  Island  Sound,  and  Matlacha  Pass  create  a  strong  seaward 
surface  residual  flow  at  the  lower  Charlotte  Harbor. 

The  offshore  residual  flow  in  the  Gulf  of  Mexico  is  northward.  The  strong  outflows 
from  the  Boca  Grande  Pass  and  San  Carlos  Pass  create  the  clockwise  gyres  near  the  surface 
in  the  Gulf  of  Mexico. 

The  residual  salinity  has  a  uniform  vertical  and  horizontal  distribution  in  the  lower 
Charlotte  Harbor,  while  the  salinity  has  horizontal  inclination  across  the  channel  and 
vertically  stratification  in  the  upper  Charlotte  Harbor.  In  upper  Charlotte  Harbor,  the  surface 


107 
salinity  along  the  right  side  shoreline  has  greater  value  than  that  along  the  left  side  shoreline 
because  of  the  residual  surface  flow  pattern.  The  residual  salinity  field  shows  a  vertical 
stratification  of  2  ppt  at  the  Peace  River  mouth  since  the  strong  river  discharge  and  mixed 
tide  from  Gulf  of  Mexico  create  two  layered  flow  (seaward  flow  at  surface  and  landward 
flow  at  bottom)  in  this  area.  A  two  layered  flow  and  salinity  stratification  is  often  presented 
at  the  Charlotte  Harbor  estuarine  system.  The  significant  vertical  stratification  can  suppress 
vertical  mixing  and  lead  to  the  formation  of  pycnocline  and  hypoxia  in  the  upper  Charlotte 
Harbor  (Stoker,  1992;  CDM,  1998).  Therefore,  understanding  of  the  density-driven 
circulation  is  critical  to  the  management  of  Charlotte  Harbor  estuarine  system. 


108 


100 


100  r 


■100 


190 


Myakka  River  at  El  Jobean  (site  1)   July,  1986 


Julian  Day 
Peace  River  at  Punta  Gorda  (site  2)    July,  1 986 


200 


210 


Julian  Day 
Charlotte  Harbor  at  Bokeelia  (site  3)    July,  1 986 


IUU 

? 

o 

50 

i 

<D 
| 

0 

i- 
CO 

-50 

'. 

210 


Julian  Day 

Figure  5.10  Comparison  between  simulated  and  measured  water  level  in  July  1986. 


109 


Gulf  of  Mexicoat  Cayo  Costa  near  Boca  Grande  (site  5)   July,  1 986 


100  r 


-100 


100  r 


Julian  Day 
Pine  Island  Sound  near  Captiva  (site  6)   July,  1 986 


200. 


Julian  Day 
B.  Fort  Myers  Beach  (site  8)    July,  1 986 


210 


190 

Figure  5.10  continued 


ulian  Day 


110 


Site  1 


Site  2 


>- 
a  , 

T105 

to' 

■D 

c 

fto1 

o 

a 


12  3  4 

Frequency  (cycle/day) 


Site  6 


Simulated 
Measured 


Frequency  (cycle/day) 


||Q 
o 

Q. 


710- 


5iry 
f 

C103 

0) 

TJ 

|io2 

w 

|10° 

o 

CL 


2  3  4 

Frequency  (cycle/day) 


Site  5 


Frequency  (cycle/day) 


Site  8 


Simulated 
Measured 


12  3  4 

Frequency  (cycle/day) 


Figure  5.1 1  Comparison  between  simulated  and  measure  spectra  of  water  level  in  July 
1986. 


Simulated  Velocity 


111 


Measured  Velocity 


m161910(SI-1),      July  1986 


o 

<v 


o 
o 


I   I    I    I    I   I 


J l_ 


10  11  12  13 


14 


15  16 


o 


o 

o 


W^w 


y/Mi 


WYV 


LP 


-i I I L. 


9  10  11  12  13  14  15  16 

Figure  5.12  Comparison  between  simulated  and  measured  velocity  data  in  July  1986. 


112 


Measured  Velocity 


m809110(SI-2),    JULY  1986 


o 

I 

o 

o 


Simulated  Velocity 


o 

o 


Figure  5.12  continued. 


Measured  Velocity 


m810110(SI-3),      July  1986 


o 

111 

i 

o 


I l_J I ^_l l_-l 1 L_J- 


J 


10 


11 


12 


13 


14 


15 


16 


Simulated  Velocity 


o 

0) 

I 

o 

o 


_J I I I       I 1 1 L- 

10  11 


12 


13 


14 


15 


16 


Figure  5.12  continued. 


113 


114 


Measured  Velocity 


S 

E 
o 

o 


m149614(SI-6),      July  1986 


i '   i    i   i i i i i i i i — i — i — i — i — i — i — i — i — i — i 

9  10  11  12  13  14  15  16 


Simulated  Velocity 


1 


\V\ 


- 


i« M Ill 


ftfMi 


i  i  i  i  i  i 


■  ....  i  .  . i  i  i  i  i  i  i  i 


9  10  11  12  13  14  15  16 


Figure  5.12  continued. 


115 


SI-1  station,      July  1986 


SI-2  station,      July  1986 


35 
30 
25 

P 

L 

L 

£ 

=15 
s 

10 

5 


35 
30 

25 

*■» 
D. 
320 

=  15 

10 

5 

9 


simulated  salinity 
measured  salinity 


f90  195  .    ..  200 

Julian  Day 


205 


SI-3  station,      July  1986 


; 

H^^r.^y^r^^X'yW^''%-A 

- 

- 

simulated  salinity 

measured  salinity 

.    i 

90  195  .    ,.  200 

Julian  Day 


205 


35 


35 
30 
25  h 


Q. 
S20 

£15 

s 

10 


5  - 
9 


90 


simulated  salinity 
measured  salinity 


1 95  200  205 

Julian  Day 
S 1-6  station,      July  1986 


simulated  salinity 
measured  salinity 


195  ,    ..  200 

Julian  Day 


205 


35 
30 

25 

Q. 

S20 

£ 

£15 
10 

5 

9 


SI-7  station,      July  1986 


1  i  ill  ;i."iiV:c 


i    nihil! 


\m\i 


>j  ys.-^ 


90 


simulated  salinity 
measured  salinity 


195  .   .   .         200 

Julain  Day 


205 


35 


SI-8  station,      July  1986 


30  - 

25 
C 
a. 

81 

10 


simulated  salinity 
measured  salinity 


90 


195  ,    ..  200 

Julian  Day 


Figure  5.13  Comparison  between  simulated  and  measured  salinity  in  July  1986. 


205 


116 


l  Vw 


Low  Slack  Tide 
8/06/86  10:00 

1  m/sec 


Flood  Tide 
8/06/86  13:00 


!»>»-•"., 


/WW-'.'lf"':-* 


j^.^ 


High  Slack  Tide 
8/06/86  16:00 


•••  '-••'•-•-'■■••'•••••••  '•■y/4/>y^y^k 

••.••.•'.'•.'/.v:.;;^:$^ 


Ebb  Tide 
8/06/86  19:00 


I^r< 


'•.\\VWSavV 


Figure  5.14  Typical  flow  pattern  of  Charlotte  Harbor  estuarine  system  during  one  tidal 
cycle  for  August  6,  1986. 


117 


I 


& 


X. 


V;    v 


<A  fe*d 


Residual  Flow 
at  Surface 

1 0  cm/sec 


Residual  Flow 
at  bottom 

1 0  cm/sec 


\x,     \t    h 


Residual  Salinity 
at  Surface 


rca* 


Residual  Salinity 
at  bottom 


,4r* 


Figure  5.15  The  29-day  residual  flow  and  salinity  for  Charlotte  Harbor  estuarine  system 
during  July  2  to  July  30,  1986 


118 
5.4  Simulations  for  2000  Hydrodynamics 

1986  hydrodynamic  simulation  of  Charlotte  Harbor  estuarine  system  was  conducted 
for  30  days  to  calibrate  the  major  model  coefficients  and  inputs.  The  field  data  did  not  cover 
the  Caloosahatchee  River  which  is  a  very  important  segment  of  the  study  area.  To 
supplement  1986  calibration  simulation,  long-term  model  calibration  was  conducted  using 
one  year  water  level  and  salinity  data  in  the  Caloosahatchee  River  during  2000.  Figure  5.16 
shows  the  salinity  measurement  stations  in  Caloosahatchee  River.  Water  levels  were 
measured  at  Fort  Myers  and  Shell  Point.  The  validated  CH3D  model  was  then  used  to  assess 
the  effects  of  Sanibel  Causeway  and  the  navigation  channel  in  San  Carlos  Bay,  and  to 
quantify  the  relationship  between  freshwater  inflow  and  spatial  and  temporal  salinity 
distribution  in  the  Caloosahatchee  River. 
5.4.1  Sensitivity  and  Calibration  Simulations 

Several  simulations  were  performed  in  order  to  test  the  effects  of  grid  configuration, 
bottom  roughness,  bathymetry,  and  salinity  advection  scheme  on  salinity  distribution  in  the 
Caloosahatchee  River.  To  study  how  the  horizontal  grid  resolution  affects  simulated 
circulation  and  transport  within  the  estuarine  system,  two  2000  simulations  are  performed 
using  two  different  grids.  Figure  5.17  shows  the  comparison  of  the  coarse  grid  (71x92)  and 
the  fine  grid  (92x  1 29).  The  fine  grid  has  about  two  times  finer  resolution  than  the  coarse  grid 
in  San  Carlos  Bay  and  near  the  Caloosahatchee  River  mouth.  Table  5.10  shows  that  the  two 
different  horizontal  grids  affected  the  salinity  less  significantly  than  the  water  level.  The 
geometry  and  bathymetry  at  the  river  mouth  play  a  very  important  role  for  water  circulation 
inside  the  river.  The  coarse  grid  tends  to  produce  more  error  in  water  level  simulation  than 
fine  grid  because  the  coarse  grid  does  not  accurately  to  represent  the  islands  and  navigation 


119 
channels  in  the  San  Carlos  Bay  and  Caloosahatchee  River  mouth.  The  fine  grid  (92x129) 
is  used  for  all  simulations  for  Charlotte  Harbor  estuarine  system. 


o 
o 
o 
o 

CD 
CD 

CM 


T 


BR31r  ^#^ 


§ityf  Fort  Myers 


JKTSF*** 


1 


J L 


h.-B-K^   I       I       I 


400000 


410000        420000 

Easting  (m) 


S79 


-L 


430000 


Figure  5.16  Locations  of  the  available  2000  water  level  and  salinity  measured  stations  at 
Caloosahatchee  River  operated  by  SFWMD. 


120 


depth 
(cm) 

■i   1000 


900 
800 
700 
600 
500 
400 
300 
200 
100 
0 


t 


depth 
(cm) 

1000 


900 
800 
700 
600 
500 
400 
300 
200 
100 
0 


•A> 


Figure  5.17  The  comparison  of  the  coarse  grid  (71x92)  and  the  fine  grid  (92x129)  for 
Charlotte  Harbor  estuarine  system. 


121 

Table  5.10  The  effect  of  horizontal  grid  resolution,  on  the  accuracy  of  simulated  water  level 
and  salinity.  Values  shown  are  average  RMS  errors  for  2000  calibration  at  all  available 
stations.  Values  shown  in  parenthesis  are  %  RMS  error  normalized  by  maximum  values. 

Variables  Station  71x92  92  x  129 


S79 

S 
B 

1.69(4.62) 
2.54(6.93) 

1.46(3.98) 
1.84(5.03) 

Salinity 
(Ppt) 

BR31 

S 
B 

1.17(3.20) 
1.89(5.16) 

1.55(4.22) 
2.06(5.61) 

Fort  Myers 

S 
B 

1.55(4.23) 
2.07(5.63) 

1.85(5.03) 
2.63(7.18) 

Shell  Point 

S 
B 

3.78(10.29) 
3.37(9.17) 

3.76(10.25) 
3.53(9.64) 

Sanibel 

S 
B 

1.85(5.05) 
2.05(5.59) 

1.94(5.28) 
2.00(5.45) 

average 

2.20(5.99) 

2.26(6.17) 

Water  level 

Fort  Myers 

9.08(5.44) 

5.51(3.30) 

(cm) 

Shell  Point 

6.92(4.83) 

4.25(2.96) 

average 

8.00(5.14) 

4.58(3.13) 

Vertical  salinity  stratification  in  the  Upper  Charlotte  Harbor  and  Caloosahatchee 
River  is  a  common  seasonal  occurrence  (Environmental  Quality  Laboratory,  Inc.,  1979).  In 
high  river  inflow  events,  a  stable  vertical  salinity  gradient  is  created  which  suppresses 
vertical  mixing  unless  there  are  sufficient  mixing  by  wind  or  tide.  Therefore,  the  vertical 
grid  resolution  is  an  important  factor  to  reproduce  vertical  salinity  distribution  which  is  a  one 
of  major  cause  effects  of  hypoxia  in  upper  Charlotte  Harbor.  To  study  the  effect  of  varying 
the  number  of  vertical  layers  used  by  CH3D  model,  2000  simulations  were  performed  with 
4  and  8  vertical  layers  (Table  5.11).  Since  a  one  year  simulation  of  hydrodynamics,  sediment 
transport,  and  water  quality  took  several  days,  the  simulation  with  more  than  8  vertical  layers 
which  required  the  additional  computational  time,  is  not  feasible  for  this  study.  While  the 
simulated  water  level  RMS  error  changes  little,  the  averaged  simulated  salinity  RMS  error 


122 


improves  about  1  ppt  with  8  layer  simulation.  The  largest  improvement  in  simulated  salinity 

was  achieved  at  Shell  Point  station  with  3  ppt  using  8  vertical  layers.  To  reproduce  vertical 

stratification  for  salinity  and  water  quality  simulation,  eight  layer  simulations  are  deemed 

appropriate  for  the  Charlotte  Harbor  estuarine  system. 

Table  5.1 1  The  effect  of  vertical  grid  resolution,  on  the  accuracy  of  simulated  water  level 
and  salinity.  Values  shown  are  average  RMS  errors  for  2000  calibration  at  all  available 
stations.  Values  shown  in  parenthesis  are  %  RMS  error  normalized  by  maximum  values. 

8  layers 


Variables 


Station 


4  layers 


S79 

S 
B 

1.85(5.08) 
2.09(5.73) 

1.46(3.98) 
1.84(5.03) 

Salinity 
(Ppt) 

BR31 

S 
B 

1.86(5.11) 
2.37(6.51) 

1.55(4.22) 
2.06(5.61) 

Fort  Myers 

S 
B 

2.34(6.43) 
2.30(6.31) 

1.85(5.03) 
2.63(7.18) 

Shell  Point 

S 
B 

6.82(18.70) 
5.31(14.55) 

3.76(10.25) 
3.53(9.64) 

Sanibel 

S 
B 

2.48(6.80) 
4.02(11.02) 

1.94(5.28) 
2.00(5.45) 

average 

3.14(8.63) 

2.26(6.17) 

Water  level 

Fort  Myers 

4.50(2.70) 

5.51(3.30) 

(cm) 

Shell  Point 

4.42(3.08) 

4.25(2.96) 

average 

4.46(2.89) 

4.58(3.13) 

In  Caloosahatchee  River,  water  level  and  salinity  are  very  sensitive  with  the  varying 
bottom  roughness  while  the  varying  bottom  roughness  does  not  affect  the  simulated 
circulation  and  transport  significantly  in  the  estuary  in  July  1986.  However,  several  2000 
simulations  were  performed  using  different  coefficient  value  to  determined  the  effect  of 
varying  the  bottom  roughness  on  the  accuracy  of  the  simulation  circulation  and  transport. 
(Table  5.12).  With  increasing  bottom  roughness,  accuracy  of  water  level  is  improved  while 
that  for  salinity  is  worse.  A  constant  bottom  roughness  of  0.3  cm  is  used  for  all  simulations. 


123 

Table  5.12  The  effect  of  varying  bottom  roughness,  z0,  on  the  accuracy  of  simulated  water 
level  and  salinity  in  2000.  Values  shown  are  average  RMS  errors  at  all  data  stations. 

Bottom  roughness  (cm)         Water  level  (cm) Salinity  (ppt) 

0.2                                    4.94                                2.28 
0.3                                   4.87                               2.29 
04 4J38 231 

Model  sensitivity  study  was  conducted  to  investigate  the  relative  accuracy  of  the 
salinity  advection  schemes  including  upwind,  QUICKEST,  and  Ultimate  QUICKEST 
methods.  Time  series  comparisons  of  measured  and  simulated  surface  salinity  using  three 
advection  schemes  at  three  stations,  Shell  Point,  Fort  Myers,  andBR31,  are  shown  in  Figures 
5.18  to  5.20.  The  salinity  comparison  at  Shell  point  does  not  show  much  difference  for  all 
three  schemes,  while  at  Fort  Myers  and  BR31,  the  simulated  salinity  with  upwind  scheme 
is  much  higher  than  those  for  the  other  advection  schemes  and  measured  salinity.  The 
advection  scheme  plays  a  very  important  role  for  salinity  simulation  at  cells  inside  river 
where  there  is  usually  very  strong  salinity  gradient  from  the  river  mouth  at  Shell  Point  to  the 
upstream.  Lower  order  advection  scheme  such  as  upwind  scheme  tends  to  produce  more 
error  in  salinity  simulation  because  of  its  inherently  high  numerical  diffusion.  A  series  of 
vertical-longitudinal  salinity  profiles  along  the  axis  of  the  river  during  slack  water  is  shown 
in  Figure  5.21  to  compare  salinity  distribution  along  the  river  simulated  by  various  advection 
schemes.  The  salinity  distribution  obtained  with  the  QUICKEST  method  shows  similar 
pattern  with  that  for  Ultimate  QUICKEST  method.  The  upstream  salinity  near  BR3 1  shows 
much  higher  value  with  upwind  scheme  because  salinity  is  quickly  diffused  from  river  mouth 
to  the  upstream.  With  the  upwind  scheme,  it  is  difficult  to  reproduce  the  vertical  salinity 
stratification  which  often  occurs  during  high  river  flow  period,  while  the  other  advection 
schemes  produced  the  stratification. 


124 


Ultimate  QUICKEST  Scheme 


200  250 

QUICKEST  Scheme 


D. 

a. 

>. 

4-1 

c 


200 
Upwind  Scheme 

35  r 


250 


simulated  salinity 
measured  salinity 


300 


300 


200 


250 


300 


350 


350 


350 


Julian  Day 

Figure  5.18  A  comparison  between  simulated  and  measured  salinity  at  Shell  Point  using 
Ultimate  QUICKEST,  QUICKEST,  and  upwind  advection  schemes. 


125 


Ultimate  QUICKEST  Scheme 


35 

30 

a-  25 

Q. 

S    20 

i    15 
5 

w    10 

5 
0 


200  250 

QUICKEST  Scheme 


35 

30 

S?   25 

Q. 

3   20 

I    15 

CO 

w    10 

5 
0 


200 

Upwind  Scheme 


250 


^A^^ 


- 

simulated  salinity 

measured  salinity 

300 


300 


350 


350 


200 


250 


300 


350 


Julian  Day 

Figure  5.19  A  comparison  between  simulated  and  measured  salinity  at  Fort  Myers  using 
Ultimate  QUICKEST,  QUICKEST,  and  upwind  advection  schemes. 


126 


Ultimate  QUICKEST  Scheme 


20 


-,    15 

a 

S 

.IT  10 

c 

TO 


200  250 

QUICKEST  Scheme 


20 


^    15 

a 
a. 

£•   10 

c 

nj 


0 


200 

Upwind  Scheme 


250 


~    15  - 


Q. 

a 


£*    10 


200 


250 


simulated  salinity 
measured  salinity 


300 


300 


350 


350 


300 


350 


Julian  Day 

Figure  5.20  A  comparison  between  simulated  and  measured  salinity  at  BR31  using 
Ultimate  QUICKEST,  QUICKEST,  and  upwind  advection  schemes. 


127 


Current  Time  :  9/  7/2000  1 9:00 

100  Shell  Fort 

Point  Myers 


-40000 


BR31 


-30000         -20000         -10000 

Distance  froma  S79  (cm) 


Salinity 
S79    <PP*> 


100  Shell 
Point 


Fort 
Myers 


BR31 


-40000 


-30000  -20000  -10000 

Distance  from  S79  (cm) 


Salinity 
S79    (ppt) 


BR31 


-40000 


10000 


Salinity 
S79    (Ppt) 


-30000  -20000 

Distance  from  S79  (cm) 

Figure  5.21  Simulated  longitudinal-vertical  salinity  along  the  Caloosahatchee  River  at 
slack  water  before  flood  on  September  7,  2000. 


128 

Table  5.13  shows  the  effect  of  salinity  advection  scheme  on  the  accuracy  of  simulated 

water  level  and  salinity.  The  results  show  slight  difference  between  water  level  and  salinity 

simulated  with  the  Ultimate  QUICKEST  scheme  and  the  QUICKEST  scheme,  while  the 

simulated  salinity  RMS   errors  with   upwind  scheme  is   much  greater.      Maximum 

improvement  in  simulated  salinity  is  achieved  when  the  QUICKEST  scheme  is  used  because 

most  of  the  model  parameters  and  boundary  conditions  are  calibrated  using  the  QUICKEST 

scheme.    However,  because  of  the  higher  cost  of  model  simulations  using  the  Ultimate 

QUICKEST  scheme,  coupled  with  only  a  marginal  improvement  in  the  simulated  results,  the 

QUICKEST  scheme  is  deemed  appropriate  for  this  study. 

Table  5.13  The  effect  of  varying  salinity  advection  scheme  on  the  accuracy  of  simulated 
water  level  and  salinity  in  2000.  Values  shown  are  average  RMS  errors  at  all  data  stations. 

Advection  scheme Water  level  (cm) Salinity  (ppt) 

2.35 
2.34 
3/73 

Beside  adjusting  boundary  condition  and  model  coefficients,  it  is  also  interesting  to 
see  how  varying  the  model  grid  and  bathymetry  affects  simulated  circulation  and  transport 
within  the  estuary.  Because  flow  through  the  Caloosahatchee  River  mouth  is  crucial  to  the 
water  level  and  salinity  inside  the  river  basin,  it  is  useful  to  perform  a  numerical  simulation 
to  study  how  the  flow  through  the  river  mouth  is  affected  by  its  cross-sectional  area.  As  was 
discussed  in  the  previous  section,  the  Charlotte  Harbor  grid  bathymetry  was  developed  by 
interpolating  measured  bathymetry  onto  the  entire  grid  followed  by  a  simple  smoothing 
scheme.  Although  the  bathymetry  in  upper  Charlotte  Harbor  and  inside  Caloosahatchee 
River  were  updated  by  measured  data  by  SWFWMD  and  SFWMD,  the  resolution  of 
bathymetry  in  San  Carlos  Bay  area,  including  the  Caloosahatchee  River  mouth,  is  very  coarse 


Ultimate  QUICKEST 

5.56 

QUICKEST 

5.63 

Upwind 

5.36 

129 
(about  300  m  resolution).  To  study  how  the  accuracy  of  grid  bathymetry  affects  simulated 
the  circulation  and  transport,  several  2000  simulations  were  performed  using  grid  systems 
with  modified  bathymetry  data.  The  first  simulation  used  a  modified  grid  bathymetry  which 
had  a  minimum  depth  of  2.5  m  (1.5  m  for  baseline  simulation).  This  simulation  was 
performed  to  determine  how  important  the  "shallowness"  of  the  river  is  to  the  circulation  and 
transport.  The  second  simulation  used  a  bathymetry  that  included  an  artificially  dredged 
channel  through  out  the  entire  river.  The  navigation  channel  included  in  the  grid  bathymetry 
at  a  depth  of  3.5  m  (NAVD88).  Since  the  grid  system  is  too  coarse  to  adequately  resolve  the 
navigation  channel,  forcing  a  channel  into  the  grid  system  resulted  in  over  estimation  of  the 
cross-sectional  areas  within  the  river.  Table  5.14  compares  the  results  of  the  year-long 
simulations  performed  using  the  original  and  two  modified  grid  bathymetries.  While  forcing 
an  deeper  navigation  channel  into  the  grid  system  had  little  effect,  imposing  a  2.5  m 
minimum  cell  depth  slightly  improved  simulated  salinity  and  greatly  worsened  water  level 
through  the  Caloosahatchee  River.  Since  neither  modified  grid  system  improved  the  2000 
simulation,  the  standard  bathymetry  is  used  for  all  subsequent  simulations. 

Table  5. 14  The  effect  of  modifying  bathymetry  on  the  accuracy  of  simulated  water  level  and 
salinity  in  2000.  Values  shown  are  average  RMS  errors  at  all  data  stations. 


Simulation  description 

Water  level  (cm) 

Salinity  (ppt) 

Standard  bathymetry 

4.88 

2.29 

Minimum  depth  of  2.5  m 

6.14 

1.90 

Deeper  navigation 
channel 

4.90 

2.32 

Overall,  the  model  is  able  to  simulate  water  level  and  salinity  within  10  %  normalized 
RMS  error  with  the  maximum  measured  values.  This  indicates  that  the  hydrodynamic  model 
reproduces  the  basic  circulation  and  salinity  transport  of  the  Caloosahatchee  River  so  that 


130 
this  model  can  be  used  to  simulate  water  quality  processes. 
5.4.1  Results  of  the  2000  Simulation 

This  section  summarizes  the  calibration  of  the  long  term  Charlotte  Harbor  circulation 

and  transport  model  using  the  field  data  collected  in  January  to  December  2000.  In  addition, 

the  mathematical  model  of  the  temperature  transport  with  heat  flux  model  as  surface 

boundary  condition  was  applied  to  improve  the  simulation  of  circulation  in  the  study  area. 

Based  on  the  results,  it  is  apparent  that  the  model  accurately  simulated  the  observed  water 

level,  salinity,  and  temperature  distribution  in  the  Caloosahatchee  River.    Table  5.15 

summarizes  the  boundary  conditions  and  model  parameters  for  2000  simulation. 

Table  5.15    A  summary  of  boundary  conditions  and  model  parameters  used  in  2000 
simulation. 

Boundary  Condition  or  Model  Parameter Value 

Tidal  Forcing  Measured  at  Naples 

Wind  speed  and  direction  Measured  at  3  stations 

Fresh  Water  Discharge  Estimated  for  Estero  Bay  and  Measured  at 

Peace,  Myakka,  and  Caloosahatchee  Rivers 

Bottom  Roughness  Constant  (0.3  cm) 

Horizontal  Diffusion  Constant  (10000  cm2/s) 

Horizontal  grid  92  x  129 

Vertical  Layers  8 

Water  Level 

When  measured  water  level  at  an  open  sea  is  used  as  a  tidal  boundary  condition,  it 
is  necessary  to  unify  datum  level  with  bathymetry  and  water  level  measured  in  the  estuary. 
For  the  2000  simulation  of  Charlotte  Harbor,  the  datum  level  for  all  measured  water  levels 
and  the  bathymetry  were  converted  to  NAVD88.  Water  level  at  the  Naples  station,  which 
is  the  tidal  boundary  condition,  was  leveled  to  MLW  by  NOAA  CO-OPS  and  converted  to 


131 
NAVD88  according  to  tidal  bench  mark  at  Naples  (26°  7.8'N,  81°  48.4W). 

Calculated  RMS  errors  between  simulated  and  measured  water  level  for  2000 

simulations  are  shown  in  Table  5.16.  The  normalized  RMS  errors  are  less  than  3%  at  all 

available  water  level  stations,  demonstrating  the  model's  ability  to  accurately  reproduce 

surface  elevation  in  the  system. 

Table  5.16  Calculated  RMS  errors  between  simulated  and  measured  water  level  for  2000 
simulation 


Station  Name 

RMS 

!  error 

(cm) 

Max 

imum  Range 
(cm) 

%  RMS  error 

Fort  Myers 

5.51 

166.80 

3.30 

Shell  Point 

4.25 

143.26 

2.96 

Average 

4.88 

155.03 

3.13 

Figure  5.22  shows  a  year  long  and  20  days'  comparison  between  simulated  and 
measured  water  level  at  Shell  point  and  Fort  Myers.  Although  the  normalized  RMS  errors 
are  very  small,  the  model  constantly  overestimated  at  Fort  Myers  and  underestimated  at  Shell 
Point  station.  A  probable  sources  of  these  differences  could  be  the  bathymetry,  grid 
resolution,  bottom  roughness,  and  open  boundary  condition.  In  the  Caloosahatchee  River, 
there  is  a  very  narrow  navigation  channel  which  the  current  grid  system  could  not  resolved. 
The  bottom  roughness  is  a  very  sensitive  parameter,  specially  in  the  river  environment,  as 
shown  in  previous  calibration.  This  study  uses  a  constant  bottom  roughness  of  0.3  value,  but 
there  is  no  direct  measurement  of  the  bottom  roughness  for  this  estuarine  system.  This  might 
cause  some  error  in  model  simulation.  In  this  study,  CH3D  does  not  include  calculation  of 
flooding/drying  cell.  Therefore,  this  study  uses  a  minimum  depth  of  1.5  m  which  could  not 
resolve  any  shallow  water  below  this  depth.  Once  again,  it  has  been  demonstrated  that  model 
is  sensitive  to  the  bathymetry  and  grid  resolution,  and  the  model  accuracy  will  improve  if 


132 
more  accurate  bathymetry  and  grid  are  used. 

Salinity  and  Temperature 

The  circulation  in  Charlotte  Harbor  estuarine  system  is  driven  primarily  by  the  mixed 
(diurnal  and  semi-diurnal)  tides  from  the  Gulf  of  Mexico,  as  well  as  by  wind  and  density 
gradient.  During  periods  of  high  freshwater  inflow  from  the  rivers,  significant  vertical 
salinity  stratification  can  be  found  in  the  Upper  Charlotte  Harbor  and  Caloosahatchee  River. 
A  two  layered  flow  and  salinity  structure  characteristic  of  the  "classic  estuarine  circulation" 
is  often  present  (Sheng,  1998).  Therefore,  understanding  of  vertical  salinity  distribution  is 
a  very  important  part  of  hydrodynamic  and  transport  simulations.  The  salinity  is  measured 
at  two  locations  in  the  water  column  referred  to  as  "upper"  and  "lower".  The  vertical 
positions  of  the  salinity  measurements  taken  by  SFWMD  are  given  in  Table  5.17.  For  this 
study,  eight  vertical  layers  are  used  by  the  CH3D  model.  The  simulated  salinity  values  at 
eight  vertical  layers  are  interpolated  vertically  to  allow  comparison  of  salinity  at  the  exact 
location  of  the  measured  stations. 

Table  5.17  presents  the  maximum  and  minimum  measured  salinity,  the  RMS  error, 
and  the  normalized  RMS  error  (with  respect  to  maximum  salinity)  for  all  stations  during 
2000.  The  results  show  the  model's  ability  to  simulate  salinity  is  within  7%  error  except  at 
the  Shell  Point  station.  The  Shell  Point  station  is  located  at  the  Caloosahatchee  River  mouth, 
which  is  very  sensitive  to  bathymetry  and  grid  resolution.  As  mentioned  before,  to  improve 
salinity  at  this  point,  finer  grid  resolution  and  bathymetry  data  will  be  needed.  It  should  be 
pointed  out  that  %  RMS  error  is  not  a  good  measure  of  model  accuracy.  At  S79  and  BR3 1, 
although  the  %  RMS  error  is  low,  the  actual  error  is  quite  significant. 

Figure  5.23  to  5.27  show  the  near-bottom  and  near-surface  simulated  and  measured 


133 
salinity  at  SFWMD  stations  in  Caloosahatchee  River.     The  results  show  reasonable 

agreement  with  measured  data  for  both  wet  season  and  dry  season  except  at  S-79  and  BR3 1 . 

S-79  and  BR31  stations  are  directly  affected  by  the  river  boundary  condition  because  these 

two  stations  are  located  within  a  few  grid  cells  from  S-79.  The  water  level  and  discharge  at 

the  Lock  and  S-79  is  controlled  by  8  tainter  gates  and  2  sector  gates  whereas  the  river 

boundary  condition  in  CH3D  is  specified  by  the  averaged  flow  rate.  At  these  two  stations, 

there  is  only  one  grid  cell  across  the  relatively  narrow  width.  Therefore,  the  current  boundary 

condition  and  grid  resolution  at  S-79  and  BR31  stations  could  not  resolve  the  detail 

characteristics  of  circulation  and  salinity  transport.  Furthermore,  the  bathymetry  in  this  area 

was  produced  from  only  a  few  cross  sectional  bathymetry  data  provided  by  SFWMD. 

Table  5.17  Calculated  RMS  errors  between  simulated  and  measured  salinity  for  2000 
simulation 


Station 
Name 

Mean 
depth 
(cm) 

Layers 

Location 
from 

bottom 
(cm) 

RMS 
error 
(PPt) 

Minimum 

Salinity 

(PPt) 

Maximum 

Salinity 

(PPt) 

% 
RMS 
error 

S-79 

591 

upper 

298 

1.49 

0.3 

13.85 

4.07 

lower 

103 

1.88 

0.3 

14.55 

5.11 

BR31 

678 

upper 

386 

1.60 

0.3 

13.96 

4.36 

lower 

251 

2.11 

0.3 

15.51 

5.76 

Fort  Myers 

261 

upper 

227 

1.84 

0.3 

18.66 

5.01 

lower 

101 

2.56 

0.3 

20.83 

6.98 

Shell  Point 

304 

upper 

209 

3.82 

2.4 

34.50 

10.41 

lower 

93 

3.57 

3.1 

36.13 

9.72 

Sanibel 

254 

upper 

135 

1.96 

11.2 

36.69 

5.34 

lower 

19 

2.05 

14.3 

36.45 

5.59 

Average 

2.29 

6.23 

Seasonal  salinity  patterns  occurred  in  response  to  the  variation  in  volume  of 


134 
freshwater  inflow.  The  highest  measured  salinity  occurred  during  an  extended  period  of  low 
flow  in  December  2000.  During  the  wet  season,  the  fresh  water  from  S-79  reached  into  San 
Carlos  Bay.  The  salinity  in  San  Carlos  Bay  was  directly  affected  by  the  variation  of  river 
discharge  from  S-79  station. 

Figure  5.28  shows  the  simulated  and  measured  temperature  near  the  surface  at  the 
Fort  Myers  station.  The  measured  and  simulated  temperature  show  significant  annual 
variations  in  the  Charlotte  Harbor  estuarine  system.  Water  temperature  ranges  from  an 
average  of  about  30  °C  during  the  summer  to  about  15  °C  in  December  and  January.  The 
daily  fluctuation  of  water  temperature  is  about  1  to  3  °C.  The  simulation  result  shows  very 
good  agreement  with  measured  data  over  seasonal  and  daily  scales.  The  RMS  error  is  less 
than  2.5  °C  at  the  Fort  Myers  station  which  is  about  7%  RMS  error  of  the  maximum 
temperature  of  32.20  °C. 

Flow  Patterns 

Typical  flow  pattern  (vertically  averaged  flow)  in  San  Carlos  Bay  during  flood  tide 
and  ebb  tide  on  August,  7  2000  are  shown  in  figure  5.29.  The  large  flood  and  ebb  flows 
through  San  Carlos  Pass  satisfy  the  tidal  prism  in  the  Caloosahatchee  River,  San  Carlos  Bay, 
the  lower  part  of  Matacha  Pass  and  lower  extremity  of  Pine  Island  Sound.  Due  to  Sanibel 
Causeway  and  three  navigation  channels,  the  flood  tide  is  separated  into  three  major  flows 
along  the  navigation  channel.  The  flow  through  the  right  side  of  causeway  moves  toward  the 
Caloosahatchee  River ,  the  flow  through  left  side  of  causeway  moves  to  Pine  Island  Sound, 
and  flood  tide  across  the  center  channel  of  causeway  follows  navigation  channel  toward  the 
Matlacha  Pass.  During  ebb  tide,  the  flows  are  relatively  weak  and  follow  the  opposite 
directions.  Along  the  East-West  direction  navigation  channel,  there  are  two  direction  flows: 


135 
the  eastward  flow  during  flood  tide  and  the  westward  flow  during  ebb  tide. 
Residual  Flow  and  Salinity  Patterns 

Figures  5.30  and  5.3 1  show  the  one  year  residual  flow  and  salinity  in  2000.  In  the  San 
Carlos  Bay,  the  westward  residual  flow  is  dominant  in  the  surface  and  bottom,  while  there 
is  a  relatively  week  northward  flow  along  the  left  and  right  side  of  shoreline.  The  strong 
outflows  from  the  left  side  of  San  Carlos  Pass  create  the  anti-clockwise  gyres  near  the  surface 
at  the  outside  of  Sanibel  Causeway.  This  anti-clockwise  gyre  is  associated  with  the 
clockwise  gyre  in  the  Gulf  of  Mexico  which  is  shown  in  Figure  5.15  for  the  1986  simulation. 

The  residual  salinity  has  a  relatively  uniform  vertical  and  horizontal  distribution  in 
the  San  Carlos  Bay,  while  the  salinity  has  a  strong  vertical  stratification  at  the 
Caloosahatchee  River  mouth  due  to  the  strong  river  discharge  and  mixed  tide  from  Gulf  of 
Mexico.  Therefore,  the  density-driven  circulation  is  a  very  important  factor  for 
understanding  the  circulation  near  such  large  rivers  as  Peace  and  Caloosahatchee. 


136 


at  Fort  Myers 


100 


80  85 

at  Shell  Point 


3 


85 


200 


300 

simulated 
measured 


95 


100 


> 

• 

1  ,'S^fi»*^  $2 

t                        r  ff 

t 

k 

30 

200 

300 

simulated 
measured 


95 


JuliarrDay 

Figure  5.22  Comparison  between  simulated  and  measured  water  level  for  2000 
simulation. 


100 


137 


atS79 

30  r 


25 


S  20 

E 

ra     10 
CO 


ui00 


30 
25 

1    2° 

5    15 

c 

ro     10 
CO 

5 


loo 


S79  at  Caloosahatchee  River 


20?Julian  Day 


Measured  at  near  Surface 
Simulated  at  near  Surface 


£±~ 


-*/v^  ■*■ 


w 


200,    ,.        „ 

Julian  Day 


300 

Measured  at  near  Bottom 
Simulated  at  near  Bottom 


n> 


a    - 


200,    ,. 
Julian  Day 


w~ 


w 


300 


Figure  5.23  Comparison  between  simulated  and  measured  salinity  at  S-79  in  2000. 


138 


S79  at  Caloosahatchee  River 


a  20 

c 

■     10 
CO 


^^\_ 


100 


30 
25 


§■    20 


C 
CO 


15 


10 


100 


20(3ulian  Day 


300 

Measured  at  near  Surface 
Simulated  at  near  Surface 


ZL 


/T* 


200,    „        _ 

Julian  Day 


300 

Measured  at  near  Bottom 
Simulated  at  near  Bottom 


-&. 


,-    -^T^i  ta 


200,    ,.        „ 

Julian  Day 


4* 

300 


Figure  5.24  Comparison  between  simulated  and  measured  salinity  at  BR31  in  2000. 


139 


S79  at  Caloosahatchee  River 


"100 


20l2)ulian  Day 


at  Fort  Myers 


300 

Measured  at  near  Surface 
Simulated  at  near  Surface 


200i  . 
Julian  Day 


300 

Measured  at  near  Bottom 
Simulated  at  near  Bottom 


100 


200 

Julian  Day 


300 


Figure  5.25  Comparison  between  simulated  and  measured  salinity  at  Fort  Myers  in  2000. 


140 


S79  at  Caloosahatchee  River 


Julian  Day 


at  Shell  Point 

40  r 


300 

Measured  at  near  Surface 
Simulated  at  near  Surface 


TOO 


20°,     ..         « 

Julian  Day 


300 

Measured  at  near  Bottom 
Simulated  at  near  Bottom 


Q. 

a. 

i 

_£ 


100  200,   ,.       _  300 

Julian  Day 

Figure  5.26  Comparison  between  simulated  and  measured  salinity  at  Shell  Point  in  2000. 


J41 


S79  at  Caloosahatchee  River 


a. 
a, 

5 

c 
ro 


20julian  Day 


near  Sanibel  Causeway 


300 

Measured  at  near  Surface 
Simulated  at  near  Surface 


Q. 
Q. 

5 

15 
(J) 


200.    , 

Julian  Day 


300 

Measured  at  near  Bottom 
Simulated  at  near  Bottom 


TOO 


Julian  Day 


300 


Figure  5.27  Comparison  between  simulated  and  measured  salinity  near  Sanibel 
Causeway  in  2000. 


142 


at  Fort  Myers 


100 


200 

Julian  Day 


300 


35 


at  Fort  Myers 


o 

o 


3   30 

+-• 
w 

I— 
Q) 
Q. 

£ 

0) 


25  - 


o 

TO 


?9 


Simulated 
Measured 


40 


145  150 

Julian  Day 


155 


160 


Figure  5.28  Comparison  between  simulated  and  measured  temperature  at  Fort  Myers  in 
2000. 


143 


/ 

■  «c» 

q    V* 

he  I 

"  *& 

& 

7     o>St[ 

n  'A 

1    \VW 

" 

i      .^\ 

»  •             > 

°!?>                    .           '     - 

":  IIIIU1'' 


mitt 


bo 


fe7   ,>';V   .*•  £ 

I:7 


n\, 


ho,*** 

!     >    '         O^" 


jus  p         ,.r  .  -       './ 


«■■■  /  .'■  >' / '' /y,^^p j)t)£  - 


rf^f" 


Ebb  Tide 
8/7/2000  0:0 


Figure  5.29  Typical  flow  pattern  of  San  Carlos  Bay  during  ebb  and  flood  tide  for  August, 
7  on  2000. 


144 


/  '•   K 


w 


•  "it 


I""' 

s     V  *  * 

f. 


us    / 


,  ■      y-ugf  .... \^' 


"^s 


1'  y'y/W4mf§^  *  r^ 


& 


... 


s 


Residual  surface  flow 


> 


>^d 


H 


iw 


J  J) 


Jliijf 


A 


->• •' ■■  -J3£u     ^  >•<*/  -"  •'   -'  / 


SB^-fcN' 


Residual  bottom  flow 

— > 
1 0  cm/s      fen 


n 


■*  <L:M>& 


\v  u 


1 


^:=s 


^V-,7  W 


Figure  5.30  One-year  residual  flow  in  San  Carlos  Bay  in  2000. 


145 


Figure  5.31  One  year  residual  salinity  distribution  in  San  Carlos  Bay  in  2000. 


146 
5.4.2  Applications  of  2000  Hydrodynamic  Simulations 
Hydrologic  Alterations 

Hydrologic  alterations  in  the  San  Carlos  Bay  and  Caloosahatchee  River  have  taken 
many  forms  such  as  navigation  channel  and  causeway.  It  has  been  suggested  such  large 
transportation  projects  as  the  dredging  of  the  IntraCoastal  Waterway  (ICW)  and  the 
construction  of  Sanibel  Causeway  (SC)  are  linked  to  the  decline  of  scallop  populations  in 
Pine  Island  Sound  (Estevez,  1998).  A  quantitative  assessment  of  this  suggestion  is  long 
overdue.  Moreover,  to  improve  the  natural  environment  of  this  area,  it  is  necessary  to 
quantify  the  effects  of  these  artificial  hydrologic  alterations.  In  this  study,  the  calibrated 
Charlotte  Harbor  model  is  used  to  evaluate  the  effects  of  these  hydrologic  alterations. 

Figure  5.32  shows  the  location  of  eleven  stations  which  are  selected  to  quantify  the 
hydrologic  alteration,  and  the  locations  of  Sanibel  Causeway  (A)  and  IntraCoastal  Waterway 
(B).  Three  month  simulations  from  April  9  to  July  8,  2000  were  conducted  with  two 
hydrologic  alterations  in  the  Charlotte  Harbor  estuarine  system.  To  test  these  effects,  three 
cases  were  considered:  (1)  present  condition  with  Sanibel  Causeway  and  existing  bathymetry 
(BASELINE),  (2)  hypothetical  condition  with  Sanibel  Causeway  removed  (NSC),  (3) 
hypothetical  condition  with  ICW  removed  (NICW).  Figure  5.33  shows  the  bathymetry  and 
shoreline  for  each  case. 

The  simulated  Charlotte  Harbor  circulation  in  the  presence  of  and  in  the  absence  of 
the  Sanibel  Causeway  and  IntraCoastal  Waterway  are  compared  in  terms  of  the  instantaneous 
flow  from  July  3  to  July  8,  2000.  Figures  5.34  and  5.35  show  the  comparisons  of  water  level 
and  salinity  for  each  case  at  three  selected  stations:  ST05  (Pine  Island  Sound),  ST08  (San 
Carlos  Bay),  and  ST10  (Caloosahatchee  River  mouth). 


147 
The  water  level  at  all  three  stations  show  little  difference  among  the  three  cases.  The 
salinity  at  Pine  Island  Sound  (ST05)  show  little  difference  among  the  three  cases,  because 
the  impacts  of  these  hydrologic  alterations  are  very  small  and  local.  The  results  show  that 
water  level  and  salinity  transport  are  more  affected  by  the  absence  of  Intracoastal  Waterway 
(NICW)  than  by  the  absence  of  the  Sanibel  Causeway  (NSC)  except  salinity  at  ST08  which 
is  located  near  the  Sanibel  Causeway.  The  freshwater  from  Caloosahatchee  River  and 
saltwater  from  Gulf  of  Mexico  are  exchanged  through  the  ICW.  In  the  absence  of  the  ICW, 
salinity  at  San  Carlos  Bay  (ST08)  is  increased,  while  salinity  at  Caloosahatchee  River  Mouth 
(ST  10)  is  decreased  at  both  surface  and  bottom  layers  because  of  the  reduced  flow  and 
salinity  transport  between  the  Caloosahatchee  River  and  San  Carlos  Bay. 

Calculated  water  level  RMS  differences  between  the  baseline  simulation  and  two 
alteration  cases  from  April  to  July  2000  are  shown  in  Table  5.18  for  11  selected  stations. 
The  RMS  differences  are  less  than  2  cm  at  all  selected  stations.  The  highest  difference  is 
found  at  station  10  which  is  located  at  the  Caloosahatchee  River  mouth.  Calculated  salinity 
RMS  differences  between  baseline  simulation  and  the  two  cases  during  this  period  are  shown 
in  Table  5.19.  Salinity  for  the  two  alteration  runs  did  not  show  much  difference  with  the 
baseline  simulation  except  at  station  10. 

The  29-day  residual  flow  and  salinity  patterns  in  San  Carlos  Bay  are  shown  in  Figure 
5.36  (at  a  surface)  and  Figure  5.37  (at  a  bottom)  during  June  9  to  July  8,  2000.  Although  the 
results  show  some  slight  impact  of  the  causeway  on  the  residual  flow  as  manifested  by  the 
circulation  gyres  in  the  immediate  vicinity  of  the  causeway,  there  is  no  noticeable  impact  on 
the  salinity  distribution  in  the  San  Carlos  Bay  area.  The  causeway  islands  did  not  block  the 
flow  of  saline  ocean  water  from  entering  into  the  San  Carlos  Bay  and  reaching  Pine  Island 


148 
Sound  because  the  causeway  islands  are  already  located  in  a  very  shallow  region.  While 
there  is  strong  residual  flow  along  the  IntraCoastal  Waterway  for  the  baseline  simulation,  this 
strong  residual  flow  vanished  without  this  waterway.  This  can  explain  the  effect  of  the 
absence  of  IntraCoastal  Waterway,  which  reduced  the  flow  and  salinity  transport  between 
Caloosahatchee  River  and  San  Carlos  Bay.  Overall,  the  IntraCoastal  Waterway  and  Sanibel 
Causeway  did  not  appear  to  show  noticeable  impact  on  the  flow  and  salinity  patterns  in  the 
San  Carlos  Bay  and  Pine  Island  Sound. 


o 
o 
o 
un 

CM 
OJ 

C\J 


st-06 


st-03 


st-05 A 


\ 


r~* 


J L 


J I L 


± 


J I I L 


± 


J I L 


390000 


395000 

Easting  (m) 


400000 


Figure  5.32  The  locations  of  Sanibel  Causeway  and  Intracoastal  Waterway  and  stations 
for  comparing  the  effects  of  hydrologic  alterations. 


149 


Depth 
(NAVD88  :cm) 

1500 
450 
400 
350 
300 
250 
200 
150 
100 


Depth 
(NAVD88  :cm) 

■  500 

450 
400 
350 
300 
250 
200 
150 
100 


Figure  5.33  The  comparison  of  bathymetry  and  shoreline  for  each  hydrologic  alteration  case 
scenarios  which  are  Baseline,  the  absence  of  IntraCoastal  Waterway,  and  the  absence  of 
causeway. 


150 


ST05 


E 
o 


CD 
> 
CD 


S 

CO 


£ 

> 

CD 


CD 

*-» 
CO 


2 

CO 


185 


ST08 


60 
40  E- 

20 

o  r 

-20 
-40 
-60 


185 


186 


186 


187Julian  Day388 


Julian  Day 


189 


190 


190 


BASELINE 
No  Causeway 
NolCW 


189 


190 


^JulianDay188 

Figure  5.34  The  comparisons  of  water  level  for  three  cases  at  three  selected  stations: 
ST05  (Pine  Island  Sound),  ST08  (San  Carlos  Bay),  and  ST10  (Caloosahatchee  River 
mouth). 


151 


ST05 


35 


Q. 

330 

c 

I25 

o 

o 

to 

•C20 
3 
W 


15 


BASELINE 
No  Causeway 
NolCW 


186 


Julian  Day 


ST08 


35- 


Q. 

-S30 
& 

C 

I25 

a> 
o 

CO 

"C20 
3 
3) 


190 


ST05 


35 


S30 


c 

I25 

£ 
o 
§20 1- 


15 


BASELINE 
No  Causeway 
NolCW 


186 


.     .-  rO88 

Julian  Day 


1    n        rJ88 

Julian  Day 


190 


ST08 

35 

a. 
-S30 

c 

s25 

/  v  V  v 

V    ' 

E 
0 

?20 

n 

1 

1 

1 

190 


186 


1    h       J88 
Julian  Day 


190 


186 


.    ■•       ,-J88 
Julian  Day 


190 


Figure  5.35  The  comparisons  of  surface  and  bottom  salinity  for  three  cases  at  three 
selected  stations:  ST05  (Pine  Island  Sound),  ST08  (San  Carlos  Bay),  and  ST10 
(Caloosahatchee  River  mouth). 


152 


Salinity 
(PPt) 

36 


Figure  5.36  The  comparisons  of  surface  residual  flow  and  salinity  fields  for  three  cases. 


153 


'p$r*r^&*.    '-.  ^'t^-— 


Salinity 
(PPt) 

36 


35 

34 
33 
32 
31 
30 
29 
28 

fe         27 

i*ffim  26 

Figure  5.37  The  comparisons  of  bottom  residual  flow  and  salinity  fields  for  three  cases. 


154 

Table  5.18  The  effects  of  hydrologic  alteration  on  2000  water  levels.   Values  shown  are 
average  RMS  differences  with  baseline  simulation  for  all  selected  stations. 

Station  Range  (cm)  RMS  error         RMS  error       RMS  error  for 

Number  for  NSC  forNICW  NSCICW 


1 

126.58 

0.41 

0.16 

0.38 

2 

115.75 

0.76 

0.31 

0.93 

3 

115.01 

1.11 

0.40 

1.31 

4 

115.19 

0.55 

0.44 

0.92 

5 

116.69 

0.36 

0.19 

0.49 

6 

114.70 

0.43 

0.23 

0.60 

7 

114.53 

0.55 

0.30 

0.80 

8 

112.49 

0.63 

0.33 

0.90 

9 

110.60 

0.62 

0.68 

1.15 

10 

84.55 

0.40 

1.36 

1.22 

11 

111.49 

0.61 

0.39 

0.94 

Average 

126.58 

0.58 

0.44 

0.88 

155 


Table  5.19  The  effects  of  hydrologic  alteration  on  2000  salinity.  Values  shown  are  average 
RMS  differences  with  baseline  simulation  for  all  selected  stations. 


Station 

Layers 

Range  (ppt) 

RMS  error 

RMS  error 

RMS  error  for 

Name 

for  NSC 

for  NICW 

NSCICW 

1 

surface 

19.48 

0.51 

0.21 

0.49 

bottom 

11.07 

0.32 

0.08 

0.34 

2 

surface 

18.16 

0.78 

0.20 

0.80 

bottom 

11.80 

0.24 

0.06 

0.24 

3 

surface 

19.90 

0.66 

0.33 

0.70 

bottom 

18.46 

0.39 

0.20 

0.44 

4 

surface 

28.45 

0.40 

0.41 

0.50 

bottom 

26.58 

0.32 

0.21 

0.31 

5 

surface 

14.42 

0.22 

0.09 

0.21 

bottom 

9.53 

0.25 

0.07 

0.29 

6 

surface 

16.46 

0.27 

0.14 

0.28 

bottom 

12.07 

0.23 

0.08 

0.24 

7 

surface 

20.69 

0.44 

0.23 

0.50 

bottom 

11.92 

0.25 

0.10 

0.28 

8 

surface 

20.95 

0.44 

0.43 

0.61 

bottom 

20.26 

0.42 

0.40 

0.57 

9 

surface 

30.44 

0.44 

0.71 

0.69 

bottom 

30.37 

0.33 

0.51 

0.66 

10 

surface 

30.01 

0.32 

1.62 

1.79 

bottom 

30.45 

0.38 

1.33 

1.24 

11 

surface 

20.83 

0.21 

0.69 

0.80 

bottom 

18.18 

0.33 

0.37 

0.42 

Average 

surface 

0.43 

0.46 

0.67 

bottom 

0.32 

0.31 

0.46 

156 
Freshwater  Inflow  and  Salinity  in  the  Caloosahatchee  River 

The  Caloosahatchee  River  has  been  drastically  altered  for  channelized  flood-control 
and  navigational  waterway.  These  changes  have  caused  large  fluctuations  in  freshwater 
inflow  volume,  frequency  of  inflow  events,  timing  of  discharges,  and  water  quality  in  the 
downstream  estuary  (Chamberlin  and  Doering,  1998).  Therefore,  it  is  necessary  to  quantify 
the  impacts  of  freshwater  inflow  from  S-79  on  downstream  estuarine  system. 

A  minimum  flow  is  defined  by  Ch.373.042(l)  F.S  (Florida  State  Law)  as  "the  limit 
at  which  further  withdrawals  would  be  significantly  harmful  to  the  water  resources  or 
ecology  of  the  area."  Significant  harm  is  defined  in  Chapter  40E-8  F.S.  as  "the  degree  of 
impact  requiring  more  than  two  years  for  the  water  (or  biological)  resource  to  recover". 
Establishing  quantifying  relationship  between  freshwater  inflow  and  the  temporal  and  spatial 
distribution  of  salinity  in  the  river  is  the  first  step  to  determine  MFL  condition. 

Salinity  distribution  in  San  Carlos  Bay  and  adjoining  water  directly  responds  to  the 
fresh  water  inflow  from  Caloosahatchee  River  (Sheng  and  Park,  2001).  In  this  study,  the 
calibrated  Charlotte  Harbor  model  is  used  to  evaluate  the  effects  of  the  fresh  water  inflow 
from  S79  to  the  salinity  distribution  in  Caloosahatchee  River.  Using  this  calibrated  model, 
a  series  of  vertical-longitudinal  salinity  profiles  along  the  axis  of  the  Caloosahatchee  River 
during  one  tidal  cycle  for  wet  and  dry  periods  of  river  discharge  at  S-79  in  2000  are  shown 
in  Figures  5.38  and  5.39,  respectively.  The  fresh  water  (less  than  1  ppt  salinity)  from  S-79 
reached  the  river  mouth  near  Shell  Point  in  wet  season,  while  the  fresh  water  stayed  upstream 
near  BR3 1  during  dry  season.  The  salinity  distribution  at  Caloosahatchee  River  shows  much 
difference  corresponding  to  river  discharge  from  S-79. 

To  quantify  this  relationship  for  Caloosahatchee  River,  the  locations  of  specific 


157 
salinity  values  were  calculated  during  the  2000  simulation  periods.  Figure  5.40  shows  the 
time  histories  of  the  river  discharge  rate  and  the  locations  of  1, 10,  and  20  ppt  surface  salinity 
along  the  Caloosahatchee  River.  As  the  river  discharge  varied,  the  1  ppt  location  is  varied 
from  S79  to  near  Shell  Point.  The  tidal  excursion  of  the  salinity  location  is  about  2  km  for 
lppt  salinity  and  5  km  for  20  ppt  salinity.  The  salinity  locations  result  from  a  combination 
of  river  discharge  and  tide  and  wind  driven  water  circulation. 

To  test  the  relationship  between  river  discharge  and  salinity  distribution,  the  river 
discharge  at  S79  was  reduced  50%  and  increased  50%  of  current  river  discharge  condition. 
To  remove  tidal  effect,  1-day  averaged  salinity  location  was  compared.  Figure  5.41  shows 
the  locations  of  1-day  averaged  1  ppt  salinity  location  during  the  simulation  period  in 
response  to  varying  river  discharge.  The  results  show  that  1  ppt  salinity  location  move  3.4 
km  downstream  and  8.5  km  upstream,  corresponding  to  a  50%  reduction  and  a  50%  increase 
in  river  discharge,  respectively. 

The  Caloosahatchee  MFL  rule  (SFWMD)  states  that:  "A  MFL  exceedance  occurs 
during  a  365-day  period,  when  (a)  30-day  average  salinity  concentration  exceeds  10  parts  per 
thousand  at  the  Fort  Myers  salinity  station  or  (b)  a  single  daily  averaged  salinity  exceeds  a 
concentration  of  20  parts  per  thousand  at  the  Fort  Myers  salinity  station.  Exceedance  of 
either  subsection  (a)  or  subsection  (b),  for  two  consecutive  years  are  a  violation."  Figure 
5.42  show  current  conditions  of  salinity  distribution  to  compare  MFL  rule  for 
Caloosahatchee  River  during  2000  simulation  periods.  According  to  this  result,  the  salinity 
does  not  exceed  the  Caloosahatchee  MFL  rule  during  2000  simulation  periods  except  winter 
dry  season. 

To  quantify  the  Caloosahatchee  MFL  rule  due  to  river  discharge,  the  locations  of  10 


158 
ppt  surface  salinity  values  were  calculated  during  the  2000  simulation  periods.  Figure  5.43 
shows  the  locations  of  10  ppt  surface  salinity  along  the  Caloosahatchee  River  due  to  river 
discharge  rate.  The  polynomial  regression  line  was  calculated  from  the  relationship  between 
the  locations  of  10  ppt  salinity  and  river  discharge  rate.  According  to  this  regression  line,  a 
total  river  discharge  of  15  m3/s  at  S79  produces  10  ppt  salinity  at  the  Fort  Myers  station. 

An  alternative  and  more  acceptable  approach  to  determine  if  the  MFL  has  been 
exceeded,  is  to  use  a  numerical  mass-balanced  model  in  which  flows  from  different  sources 
can  be  specified  (Edwards  et  al.,  2000).  Therefore,  the  total  river  discharge  at  S79  required 
to  produce  a  given  salinity  at  Caloosahatchee  River  can  be  estimated  with  a  calibrated  model. 
Using  the  calibrated  model,  a  diagram  describing  the  relationship  between  river  discharge 
and  the  locations  of  two  salinity  values  (1-day  averaged  20  ppt  and  30-day  averaged  10  ppt) 
in  the  river  were  generated  in  Figure  5.44.  Twelve  scenarios  with  constant  river  discharges 
of  5,  10,  15,  20,  30,  50,  75,  100,  150,  200  m3/s  at  S79  were  simulated  for  60  days.  The  60 
day  simulations  allowed  the  model  salinity  to  reach  equilibrium  condition  for  all  specified 
river  discharges.  To  compare  current  Caloosahatchee  MFL  rule  with  simulated  salinity 
results,  30-day  averaged  10  ppt  salinity  location  and  1-day  averaged  20  ppt  salinity  location 
were  calculated  for  each  case.  According  to  this  diagram,  a  total  river  discharge  of  18  m3/s 
at  S79  produces  30-day  averaged  10  ppt  salinity  at  the  Fort  Myers  station. 

To  quantify  the  relationship  between  salinity  at  Fort  Myers  station  and  river 
discharge,  1-day  and  30-days  averaged  salinity  at  Fort  Myers  station  were  plotted  vs.  the 
fresh  water  inflow  at  S79  (Figure  5.45).  The  result  show  the  minimum  flow  to  produce  a 
salinity  of  a  10  ppt  at  Fort  Myers  is  about  18  m3/s,  which  is  the  same  as  that  obtained  form 
Figure  5.44.   The  contribution  of  river  discharge  at  S79  to  spatial  and  temporal  salinity 


159 

distribution  was  successfully  quantified  with  the  integrated  model  for  Charlotte  Harbor 
estuarine  system.  This  modeling  approach  would  develop  a  management  tool  to  establish 
the  MFL  criteria,  with  long  term  salinity  and  river  discharge  data. 


rShell  Point 


5/10/2000  01:00 


-700 


40000 


-30000 


5/10/2000  04:00 


-700 


-40000 


-30000 


100 

0 

-100 

-200 


j=   -300 

8"  -400 
Q 

-500 

-600 


-700 


5/10/2000  07:00 


-30000 


Fort  Myers 


-700 


40000 


-30000 


-20000 
Distance  from  S79 


-20000 
Distance  from  S79 


-20000 
Distance  from  S79 


-20000 
Distance  from  S79 


BR31 


-10000 


-10000 


-10000 


-10000 


S-79  Salinity 
-*  (PPt) 


Figure  5.38  The  vertical-longitudinal  salinity  profiles  along  the  axis  of  the 
Caloosahatchee  River  during  wet  season  in  2000 


Salinity 
(PPt) 


Salinity 
(PPt) 

28 

24 

20 

16 

12 

8 

4 

0 


Salinity 
(PPt) 


160 


-700 


40000 


-30000 


-20000 

Distance  from  S79 


-10000 


-700 


40000 


-30000 


-20000 

Distance  from  S79 


-10000 


-40000 


-30000 


-20000 
Distance  from  S79 


-10000 


-40000 


-30000 


-20000 
Distance  from  S79 


-10000 


S-79  Salinity 
>  (PPt) 


Figure  5.39  The  vertical-longitudinal  salinity  profiles  along  the  axis  of  the 
Caloosahatchee  River  during  dry  season  in  2000 


Salinity 
(PPt) 


Salinity 
(PPt) 


Salinity 
(PPt) 


161 


E1 
ro 
.n 
o 
to 

I 


150  r 


100  - 


150 


200  250 

Julian  Day 


300 


20  ppt 


Shell 
"^oint 


Fort 


ppt  Myers 


BR31 


S79 


Figure  5.40  Time  histories  of  river  discharge  at  S79  and  the  locations  of  1,  10,  and  20  ppt 
surface  salinity  along  the  Caloosahatchee  River  during  2000  simulation  period. 


162 


0 

i- 


o    150 


> 


200 


300 


Location  of  1  ppt  surface  salinity 


45000  r 


40000  - 


£35000  r 


K30000 

E25000 
o 

^20000 
0 

C15000 

*S 

■^10000 
Q 

5000 


0 


current  condition 

50%  increased  river  discharge 

50%  reduced  river  discharge 


Julian  Day 


300 


Figure  5.41  Time  histories  of  river  discharge  at  S79  and  the  1  ppt  salinity  location  along 
Caloosahatchee  River. 


163 


£ 

Q 


45000 
40000 
35000 
"'30000 
25000 


0)20000 
o 

c 

*)j  15000 

Q 

10000 


5000  - 


1-day  averaged  surface  salinity  location  of  20  ppt 


rvn. 


Shell 
Point 


Fort 
Myers 


30-day  averaged  surface  salinity  location  of  1 0  ppt\ 


BR31 


J L 


X 


J I I L 


200 

Julian  Day 


300 


J L_  S79 


Figure  5.42  The  1-day  averaged  20  ppt  surface  salinity  location  and  30-day  averaged  10  ppt 
surface  salinity  location  during  the  2000  simulation  period. 


164 


0 20000 

O 

C  i  / 

ra  §K«-\  •     Minimum  Flow  =  15  m  /s 

■§15000 
Q 


10000H 
5000 


_L 


Shell  Point 


Location  of  10  ppt  salinity 
Polynomial  regression  line 


-Fort  Myers 


•BR31 


J I L 


_L 


J L 


S79 


0 


50 


100 


150 


200 


River  discharge  (m  /s) 


Figure  5.43  The  locations  of  10  ppt  surface  salinity  due  to  river  discharge  rate  at  S-79 
during  2000  baseline  simulation. 


165 


30  day  averaged  1 0  ppt  salinity 
1  day  averaged  20  ppt  salinity 


Shell  Point 


Fort  Myers 


■§15000 

Q 

10000 


BR31 


i       i       i       i 


I I I L 


S79 


50  100  150 

River  discharge  (m  /s) 


200 


Figure  5.44  The  relationship  between  locations  of  specific  salinity  value  vs.  river 
discharge  at  S-79. 


166 


♦ 1  -day  averaged  salinity 

B 30-day  averaged  salinity 


Minimum  Flow  =  18  m  /s 


-  »--- 


■  | <■  ■   f    ■   .|„         ■  |     ,„   ■!■   -   t   -  --t~ 


100 


150 


200 


River  discharge  (m  /s) 


Figure  5.45  The  relationship  between  salinity  at  Fort  Myers  station  vs.  river  discharge  at 
S-79 


CHAPTER  6 
APPLICATION  OF  WATER  QUALITY  MODEL 

The  water  quality  model  described  in  Chapter  4  was  applied  to  the  Charlotte  Harbor 
estuarine  system.  Model  applications  include  the  simulation  of  the  three-dimensional 
circulation,  sediment  transport  and  water  quality  processes  during  summer  1996,  the  entire 
2000.  These  simulations  provide  calibration  and  validation  of  the  Charlotte  Harbor  water 
quality  model  using  field  data  obtained  by  USGS,  SWFWMD,  and  SFWMD.  In  addition, 
model  simulation  provide  assessment  of  (1)  the  effects  of  hydrologic  alteration  of  causeway 
and  navigation  channels,  (2)  the  impact  of  river  pollutant  loading  and  its  reduction  on  the 
water  quality  in  the  Charlotte  Harbor  estuarine  system,  and  (3)  the  causes  for  hypoxia  of 
bottom  water  in  the  upper  Charlotte  Harbor. 

To  reduce  the  computational  time  necessary  to  simulate  annual  response  of  the 
Charlotte  Harbor  estuarine  system,  the  parallel  CH3D  model  developed  by  Sheng  et  al. 
(2003)  is  used.  The  detail  description  of  validation,  CPU  time,  and  speedup  of  the  parallel 
CH3D  code  are  given  in  Appendix  G. 

6.1  Forcing  Mechanism  and  Boundary  Condition  of  Circulation 

The  tidal  forcing  along  the  Gulf  of  Mexico  boundary  is  prescribed  by  the  water  level 
data  measured  at  Naples  by  NOS  (http://www.co-ops.nos.noaa.gov/data_res.html).  The 
surface  wind  boundary  condition  is  produced  by  using  the  hourly  wind  magnitude  and 
direction  data  collected  at  the  National  Data  Buoy  Center  C-MAN  stations  at  Venice  and 


167 


168 
USGS  station  at  Naples  and  Fort  Myers.  The  hourly  wind  magnitude  and  direction  are 
converted  into  EastAVest(-x)  and  North/South  (-y)  wind  velocity  components  and  then 
interpolated  onto  the  entire  computational  grid. 

Daily  river  discharges  for  Peace  River,  Shell  Creek,  Myakka  River,  Caloosahatchee 
River  (S-79  spillway,  Cape  Coral  and  Whiskey  Creek)  and  Estero  Bay  (Mullock  &  Hendry 
Creek,  Estero  River,  Spring  Creek,  and  Imperial  River)  were  measured  by  the  USGS  daily 
(Table  5.1).  Figures  6.1  to  6.3  show  the  water  level  at  the  tidal  boundary,  river  flows  at  the 
river  boundary,  wind  speed  and  direction  at  the  surface  boundary,  and  air  temperature  for  the 
air-sea  heat  flux  in  1996.  The  forcing  mechanism  and  boundary  condition  for  circulation 
of  2000  simulation  were  explained  at  Chapter  5. 


169 


1 50  200 

Julian  Day 


300  r 


Caloosahatchee  River 
Peace  River 
Shell  Creek 
Myakka  River 


150 


200 


Julian  Day 

Figure  6.1  Tidal  forcing  and  river  discharges  for  1996  simulations  of  Charlotte  Harbor. 


at  Venice 


170 


_1_ 


-J I : i_ 


150  160  170  180  190  200 

Julian  Day 


210 


220 


230 


at  Fort  Myers 


150  160  170  180  190  200  210  220  230 

Julian  Day 

Figure  6.2  Wind  velocity  for  1996  simulations  of  Charlotte  Harbor. 


171 


at  Venice 


150 


Julian  Day 


200 


at  Fort  Myers 


150 


200 


Julian  Day 

Figure  6.3  Air  temperature  for  1996  simulations  of  Charlotte  Harbor. 


172 
6.2  Initial  and  Boundary  Condition  for  the  Water  Quality  Model 

For  1996  simulations,  there  were  21  stations  sampled  monthly  for  water  quality  data 
in  1996  by  SWFWMD  and  SFWMD  (Table  6.1).  The  locations  of  the  measurement  sites  are 
shown  in  Figure  6.4.  Most  of  the  data  were  collected  from  May  to  July  1996.  For  Estero 
Bay,  there  were  14  sites  sampled  for  total  nitrogen,  total  phosphorous,  pH,  dissolved  oxygen, 
temperature,  turbidity  and  chlorophyll_a  data  in  June  1996  by  SFWMD.  The  initial  water 
column  concentrations  of  several  water  quality  parameters  are  determined  from  the  EPA 
data.  The  water  quality  data  for  May  are  used  to  produce  the  initial  condition  of  1996 
simulations.  The  water  quality  data  at  each  grid  cell  are  calculated  by  interpolation  of  the 
data  at  the  three  closest  data  stations,  with  a  weighting  function  inversely  proportional  to  the 
distance  from  each  of  the  three  stations.  Data  collected  at  station  CH001,  CH029,  and 
CH004  are  used  to  provide  river  loading  data  of  Myakka  River,  Peace  River,  and  Horse 
Greek,  respectively.  There  are  three  river  boundary  conditions  at  Caloosahatchee  River, 
which  are  S-79  spillway,  Cape  Coral,  and  Whiskey  Creek.  For  these  three  river  boundary 
conditions,  data  at  HB01,  HB03,  and  HB04  were  used.  The  water  quality  data  at  stations 
EB002,  EB004  EB01 3  and  EBO 1 2  are  used  for  four  river  boundary  conditions  of  Estero  Bay. 

For  2000  simulation,  the  initial  water  column  concentrations  of  several  water  quality 
parameters  were  determined  from  SWFWMD  and  SFWMD  data  of  January  2000.  Figure 
6.5  shows  the  locations  of  SWFWMD  (CH-001  to  CH-014)  and  SFWMD  (CES01  to  CES08) 
water  quality  monitoring  stations  (Table  6.1).  Data  collected  at  station  CH001,  CH029, 
CH004,  CES01,  CES05,  and  CES08  are  used  to  provide  river  loading  data  of  each  river 
boundary. 


173 


Table  6.1  Locations  of  water  quality  measured  stations 


station 


Latitude       Longitude 


XUTM 


YUTM 


(l,J)  in  Grid   Agency 


CH-001 

27  00  072 

82  15  108 

CH-002 

26  57  216 

82  12  300 

CH-02B 

26  58  048 

82  1 1  048 

CH-029 

27  00  588 

81  59  030 

CH-004 

26  56  396 

82  03  324 

CH-005 

26  55  558 

82  06  156 

CH-05B 

26  57  114 

82  06  330 

CH-006 

26  54  006 

82  07  090 

CH-007 

26  52  396 

82  04  072 

CH-009 

26  49  132 

82  05  294 

CH-09B 

26  53  132 

82  09  288 

CH-011 

26  44  120 

82  10  000 

CH-013 

26  41  300 

82  18  010 

CH-014 

26  39  310 

82  19  210 

HB-001 

26  41  480 

81  49  277 

HB-002 

26  40  007 

81  52  213 

HB-003 

26  38  090 

81  54  294 

HB-004 

26  33  576 

81  54  534 

HB-005 

26  30  570 

81  59  024 

HB-006 

26  29  250 

82  01  150 

HB-007 

26  32  264 

82  07  324 

CES01 

26  43  199 

81  41  233 

CES02 

26  43  354 

81  42  284 

CES03 

26  43  001 

81  45  382 

CES04 

26  40  541 

81  50  017 

CES05 

26  38118 

81  53  193 

CES06 

26  34  563 

81  54  367 

CES07 

26  31  488 

81  57  562 

CES08 

26  31  239 

82  00  312 

375674.130  2987267.910 

380063.200  2982147.510 

382419.400  2983448.440 

402367.040  2988631 .380 

394887.010  2980722.130 

390352.680  2979406.960 

389904.080  2981718.790 

388859.690  2975881.310 

393859.500  2973345.640 

391542.120  2966995.710 

384983.960  2974439.100 

383975.740  2957800.680 

370636.040  2952944.680 

368386.980  2949305.790 

417982.560  2953104.130 

413179.170  2949843.840 

409616.420  2946422.590 

408897.610  2938704.790 

401965.780  2933186.980 

398261.910  2930385.140 

387873.180  2936041.360 

431400.582  2955824.308 

429607.466  2956326.423 

424351.809  2955279.834 

417059.745  2951448.633 

411552.356  2946470.516 

409380.629  2940485.917 

403804.420  2934742.217 

399508.606  2934006.048 


(38,123) 

SWFWMD 

(37,120) 

SWFWMD 

(42,119) 

SWFWMD 

(71,117) 

SWFWMD 

(60,116) 

SWFWMD 

(53,116) 

SWFWMD 

(54,118) 

SWFWMD 

(47,113) 

SWFWMD 

(57,111) 

SWFWMD 

(47,106) 

SWFWMD 

(38,113) 

SWFWMD 

(29,97) 

SWFWMD 

(9,98) 

SWFWMD 

(7,95) 

SWFWMD 

(79,58) 

SFWMD 

(76,58) 

SFWMD 

(73,58) 

SFWMD 

(70,55) 

SFWMD 

(63,50) 

SFWMD 

(45,43) 

SFWMD 

(24,70) 

SFWMD 

(91 ,56) 

SFWMD 

(89,56) 

SFWMD 

(84,56) 

SFWMD 

(78,57) 

SFWMD 

(74,56) 

SFWMD 

(71,55) 

SFWMD 

(65,54) 

SFWMD 

(58,51) 

SFWMD 

174 
Light  and  temperature  are  the  main  limiting  factors  for  the  phytoplankton  growth  rate. 
Temperature  is  calculated  by  solving  the  heat  equation  with  the  heat  flux  model  as  a  surface 
boundary  condition,  which  was  explained  in  chapter  3  and  4.  The  input  data  for  the 
temperature  model  include  an  air  temperature  and  cloud  cover  rate.  Continuous  hourly  air 
temperature  data  collected  at  the  NOS  data  collection  platforms  (DCP)  and  stored  in  the  CO- 
OPS databases  are  available  for  the  Naples  and  Vince  stations.  Since  there  is  no  available 
data  for  cloud  cover,  a  constant  cloud  cover  of  0.2  is  used  for  all  simulations. 

To  calculate  PAR  (Photosynthetically  Active  Radiance)  in  the  light  attenuation 
model,  the  light  intensity  at  the  surface  is  used  as  the  surface  boundary  condition.  The  light 
intensity  data  (Langleys/day)  are  converted  from  the  global  and  diffuse  horizontal  solar 
irradiance  data  (W/m2)  which  were  processed  at  the  National  Renewable  Energy  Laboratory 
(NREL).  Figure  6.6  shows  the  light  intensity  data  used  in  the  1996  and  2000  water  quality 
model  simulations. 

A  total  of  215  sediment  grab  samples  and  28  shallow  cores  were  collected  from  the 
Charlotte  Harbor  estuarine  system  during  the  period  of  December  27,  1964,  to  January  1, 
1965  by  Huang  (1965).  Based  on  the  sediment  size  distribution,  the  entire  study  area  is 
characterized  into  five  sediment  types  (Table  6.2). 
Table  6.2  Sediment  types  for  Charlotte  Harbor  water  quality  simulations 


Type 

D50  Range         ( mm) 

Category 

Remark 

very  coarse 

D50>0.50 

5 

coarse 

0.25     >  D50  >  0.50 

4 

medium 

0.125    >D50>0.25 

3 

fine 

0.0625  >D50>  0.125 

2 

silts  or  clay 

0.0625  >  D50 

1 

Sediments  were  predominantly  sand,  with  low  amounts  of  silt,  clay,  and  organic 


175 
matter.  Silt  or  clay  deposits  were  only  observed  at  the  mouth  of  Peace  River.  This  type  of 
sediment  can  be  considered  fine  or  cohesive  sediment.  Fine  sediments  exist  in  upper 
Charlotte  Harbor,  the  southern  part  of  Pine  Island  Sound,  and  the  northern  part  of  San  Carlos 
Bay  due  to  weak  wave  actions  there.  Sediment  nutrient  analyses  were  performed  by  the 
FDEP  Sediment  Contaminant  Survey  with  data  from  33  sample  stations  from  1985  to  1989 
(Schropp,  1998).  These  data  include  organic  carbon,  total  nitrogen,  and  total  phosphorous. 
Based  on  water  quality  data,  sediment  data,  bathymetry,  and  geometry,  the  Charlotte 
Harbor  study  area  is  divided  into  15  segments  as  shown  in  Figure  6.7  for  the  water  quality 
simulations. 


176 


o 
o 
o 

O) 

cm 


C 

IE 

■c 
o 

Z  o 
o 

o 

CO 
CM 
OJ 

CM 


o 
o 
o 
o 
o 

CD 
CM 


I 


X 


CH-lp'l 

\-:.    CH-02| 


fCH-029 


V4,.  ^hdS      iKU-.n 


^  .  CH-09B*    •^r007 


At' 


% 


ak     •cfi-009 

vj  f 


•  ch-01  r 

CH-013*    (k«r€*^   J 
CH-01 4  •  W%f  &*\   %i 


HB-001    Jft^-yr- 

v.fc*" 


.  •'HB-002 


^^?or* 


fl 


<K: 


v 


1. 


i    i    i    i 


_L 


J I I L 


J I I L 


± 


i       i       i 


350000 


375000  400000 

Easting  (m) 


425000 


Figure  6.4  Locations  of  1996  water  quality  measurement  stations  operated  by  EPA 


177 


c 
o 


o 
o 
o 
m 
t-~ 

CM 


o 
o 
o 
m 
cm 

CD 
CM 


o 
o 
o 
o 
o 
<J1 

C\J 


_L 


X 


CH 


GH-02B, 


fcH-029 


X 


„    •CH-Q06 
CH-09B*  ^Hrd07 

WMEjjfaiJK       •Cf-009 


•  CH-01,1 


J_ 


CES003Ci?002 

*&£       cisob-i 

ES004 


J_ 


J L 


-L 


J L 


-L 


_ 


350000 


375000 


400000 
Easting  (m) 


425000 


Figure  6.5  Locations  of  2000  water  quality  measurement  stations  operated  by  SFWMD 
and  SWFWMD 


178 


2500  - 


2000 


£  1500 
c/) 

C 

S, 

C  1000 

+■> 
JC 

.2>  500 


— 


;  'Hi  ,!  [, 


120      140      160      180      200 

Julian  Day 


For  1996 


220  240 

For  2000 


50  100  150  200  250  300 

Julian  Day 

Figure  6.6  Light  intensity  at  water  surface  for  1996  and  2000  simulations 


350 


179 


\ 


& 


\ 


u 


"V 


fe 


i~\^--Y 


WW 


kV'    2       /-^ 


-*<w~" 


4,. 


# 


K 


# 


45*sr 


P 


*ui/ 


,/t 


X        S*^*K*f  k    «fc 


r 


Figure  6.7  Segments  for  Charlotte  Harbor  estuarine  system. 


180 
6.3  Simulations  of  Water  Quality  in  1996 

CH3D-IMS  are  used  to  simulate  the  circulation,  sediment  transport,  and  water  quality 
dynamics  of  the  Charlotte  Harbor  estuarine  system  during  May  23  to  August  21,  1996.  To 
create  an  appropriate  initial  condition  throughout  the  computational  domain,  a  spin-up 
simulation  was  executed  until  the  flow  and  salinity  field  reached  a  dynamic  steady  state.  The 
river  discharge  data  for  this  area  showed  that  the  simulation  period  can  be  divided  into  a  dry 
season  which  lasts  from  January  1  until  June  8  and  a  wet  season  which  spans  from  June  9 
until  July  22  (Sheng  and  Park,  2002).  A  30-day  spin-up  simulation  was  performed  from 
April  23  to  May  23  (dry  season)  with  all  forcing  mechanisms  (tides,  river  discharges,  wind) 
to  allow  water  level,  velocity  and  salinity  field  to  reach  dynamic  steady-state  throughout  the 
computational  domain.  Using  the  surface  elevation,  velocity,  and  salinity  at  the  end  of  the 
spin-up  simulation  as  the  initial  condition,  the  Charlotte  Harbor  circulation  and  water  quality 
models  are  then  run  from  May  23  to  August  21  with  hydrodynamic,  sediment  and  water 
quality  input  data. 
6.3.1  Calibrations 

The  calibration  of  water  quality  model  was  conducted  using  the  monthly  water 
quality  data  obtained  from  the  EPA  STORET  database.  To  compare  the  water  quality  results 
with  measured  data,  fourteen  stations  were  selected:  CH002  (near  Myakka),  CH005  (near 
Peace),  HB02,  HB03  and  HB06  (near  Caloosahatchee),  CH006,  CH007,  CH09B  and  CH009 
in  upper  Charlotte  Harbor,  CHOI  1  in  lower  Charlotte  Harbor,  HB007  in  Pine  Island  Sound, 
HB006  in  San  Carlos  Bay,  and  CHOI 3  and  CH014  in  the  open  sea. 

Several  initial  water  quality  parameters  for  the  Charlotte  Harbor  simulation  are 
obtained  from  "A  Mechanistic  Water  Quality  Model  for  the  Tidal  Peace  and  Myakka  Rivers" 


181 
(Pribble  et  al,  1997)  and  Indian  River  Lagoon  simulations  (Sheng  et  al.,  2001)  shown  in 
Chapter  4. 

In  order  to  test  the  model  response  to  variation  of  these  specific  parameters,  the 
sensitivity  test  was  conducted  for  each  water  quality  parameter  by  performing  90-day 
simulations  and  comparison  with  the  baseline  simulation.  These  simulations  used  the  same 
initial  and  boundary  conditions,  and  external  forcing  as  those  for  the  baseline  simulation. 
Therefore,  variations  in  nutrient  concentrations  between  the  sensitivity  test  and  the  baseline 
simulation  can  be  directly  related  to  variation  of  each  parameter  in  a  sensitivity  test.  Table 
6.3  shows  the  parameters  considered  in  the  sensitivity  analysis,  baseline  values  in  the 
simulations  and  the  parameter  variations  of  each  sensitivity  test.  The  tests  were  performed 
by  varying  each  parameter  within  reasonable  ranges  according  to  literature  survey.  Table  6.4 
shows  the  sensitivity  analysis  results  which  are  presented  in  terms  of  percent  RMS  difference 
in  concentration,  normalized  by  concentrations  of  each  water  quality  species  in  the  baseline 
simulation.  The  SUM  in  Table  6.4  is  linearly  averaged  value  of  RMS  errors  for  each 
sensitivity  test. 

The  results  of  the  tests  showed  the  nitrogen  half  saturation  constant  (HALN)  for 
uptake  as  the  most  sensitive  parameter  in  the  water  quality  model,  followed  by  the  maximum 
algae  growth  rate  (AGRM).  These  two  parameters  are  related  and  their  major  impact  should 
be  detected  in  the  phytoplankton  biomass.  However,  the  chlorophyll_a  concentration  is  more 
sensitive  to  ammonification  than  maximum  algae  growth  rate  or  nitrogen  half  saturation 
constant,  which  reveals  the  extension  of  nitrogen  limitation  to  phytoplankton  growth. 
Morever,  the  effect  of  reducing  the  maximum  growth  rate  (AGRM0.5)  is  more  pronounced 
in  chlorophyll_a  concentration  than  increasing  this  rate  (AGRM2.0).     Result  of  test 


182 
AGRM2.0  show  that  ammonia  nitrogen  (NH4)  and  nitrate+nitrite  (N03)  is  rapidly  uptaked 
by  phytoplankton  and  phytoplankton  is  increased  due  to  increasing  growth  rate.  If  more 
ammonia  nitrogen  and  nitrate+nitrite  are  available  in  the  system,  phytoplankton  will  increase 
more  than  that  for  this  sensitivity  test  simulation.  Related  with  algae  growth  rate  and 
nitrogen  half  saturation  constant,  algae  respiration  rate  and  algae  mortality  are  also  important 
parameters.  Therefore,  the  coefficients  related  with  algae  growth  rate  are  most  important 
factors  for  overall  water  quality  processes. 

The  third  most  important  parameter  revealed  by  the  sensitivity  test  is  an 
ammonification  rate.  Soluble  organic  nitrogen  (SON)  is  rapidly  mineralized  to  ammonium 
nitrogen.  Nitrate  (N03)  level  also  increased  due  to  nitrification.  Due  to  high  concentrations 
of  inorganic  nitrogen,  which  is  a  food  for  phytoplankton  growth,  chlorophyll_a  concentration 
increased.  On  the  opposite  side,  decreasing  an  ammonification  rate  promoted  an  increase  in 
SON,  and  decrease  in  NH4,  N03,  and  chlorophyll_a. 

The  information  obtained  from  the  sensitivity  tests  enabled  a  more  systematic  and 
efficient  calibration  of  model  coefficients  described  in  Table  4.9  of  Chapter  4.  According 
to  the  sensitivity  test,  the  six  most  important  parameters,  which  include  maximum  algae 
growth  rate,  nitrogen  half  saturation  constant,  ammonification  rate,  algae  mortality,  algae 
respiration  rate,  and  sorption/desorption  rate  for  SON  and  PON,  are  adjusted  first  for  all 
water  quality  species  as  part  of  the  systematic  calibration  procedure.  The  other  parameters 
are  adjusted  as  partially  sensitive  parameters  for  each  specific  water  quality  species.  More 
than  100  simulations  were  made  during  calibrations  which  include  all  the  sensitivity 
analyses.  During  these  simulations,  kinetic  coefficients  were  adjusted  within  accepted 
tolerances,  estimated  loads  were  reviewed  and  adjusted,  and  new  processes  were  added  or 


183 


modified  in  water  quality  model,  if  necessary.    Listed  in  Table  6.5  are  values  for  the 
calibration  parameters  described  in  Chapter  4  and  Table  4.3. 

Table  6.3  Water  quality  parameters,  baseline  values,  and  range  used  in  the  sensitivity  analysis 

Test  Run         Parameter  Literature  Baseline        Test  run  parameter 

Range               Parameter     =Baseline  parameter 
*multiplier 


AGRM2.0 
AGRM0.5 

Maximum  algae 
growth  rate 

0.2-8 

2.0  -  2.2 

2.0 
0.5 

HALN10.0 
HALN0.5 

Nitrogen  half 
saturation  rate 

1.5-400 

25 

10.0 
0.5 

HALP10.0 
HALP0.5 

Phosphorous  half 
saturation  rate 

1.0-  105 

2 

10.0 
0.5 

KAEX2.0 
KAEX0.5 

Algae  respiration 
rate 

0.02  -  0.24 

0.06 

2.0 
0.5 

KAS2.0 
KAS0.5 

Algae  mortality 

0.01-0.22 

0.07 

2.0 
0.5 

WAS  10.0 
WAS0.1 

Algae  settling 
velocity 

0.0  -  300 

10 

10.0 
0.1 

HALA2.0 
HALA0.5 

Algae  half 
saturation  rate 

200  -  2000 

200 

2.0 
0.5 

SONM2.0 
SONM0.2 

Ammonification 
rate 

0.001  -  1.0 

0.015 

2.0 
0.2 

NTTR2.0 
NITR0.5 

Nitrification  Rate 

0.004-0.11 

0.08 

2.0 
0.5 

DRON2.0 
DRON0.5 

Sorption/desorption 
rate  of  SON/PON 

0.02  -  0.08 

0.03 

2.0 
0.5 

DRAN2.0 
DRAN0.5 

Sorption/desorption 
rate  of  PIN/NH4 

0.02  -  0.08 

0.03 

2.0 
0.5 

SOPM2.0 
SOPM0.2 

Mineralization  rate 

0.001-0.6 

0.02 

2.0 
0.2 

DROP2.0 
DROP0.5 

Sorption/desorption 
rate  of  SOP/POP 

- 

0.02 

2.0 
0.5 

DRIP2.0 
DRIP0.5 

Sorption/desorption 
rate  of  SRP/PIP 

- 

0.02 

2.0 
0.5 

SODM2.0 

Sediment  oxygen 
demand 

0-  10.7 

0.5  -  2.0 

2.0 

AKD2.0 
AKD0.5 

Oxidation  rate 

0.02  -  0.6 

0.05 

2.0 

0.5 

184 


Table  6.4   Sensitivity  analysis  results  in 
quality  calibration  simulations 

RMS  d 

ifference 

w.r.t.  b 

aseline  1 

or  199 

6  water 

ChlA 

DO 

NH4 

NOX 

TKN 

P04 

PHOS 

TOC 

SUM 

BASE 

0.00 

0.00 

0.00 

0.00 

0.00 

0.00 

0.00 

0.00 

0.00 

AGRM0.5 

17.63 

4.29 

55.20 

171.6 

3.31 

25.53 

8.47 

9.51 

36.9 

AGRM2.0 

5.68 

2.54 

12.68 

12.72 

0.97 

5.00 

1.91 

1.81 

5.42 

HALN0.5 

2.03 

0.73 

6.23 

5.92 

0.47 

1.99 

0.78 

0.89 

2.38 

HALN10.0 

18.56 

4.26 

65.73 

204.0 

3.71 

37.54 

10.61 

10.99 

44.4 

HALPO.5 

0.31 

0.07 

0.35 

0.67 

0.03 

0.56 

0.10 

0.11 

0.28 

HALP10.0 

4.02 

0.77 

6.07 

13.27 

0.45 

12.03 

1.89 

1.76 

5.03 

KAEXO.5 

2.13 

0.92 

7.17 

5.24 

0.54 

2.58 

0.87 

0.92 

2.55 

KAEX2.0 

5.43 

2.08 

18.59 

17.27 

1.35 

6.87 

2.33 

2.55 

7.06 

KASSO.5 

14.75 

6.23 

2.64 

4.69 

3.47 

1.93 

1.49 

9.11 

5.54 

KASS2.0 

19.58 

5.59 

7.56 

13.49 

3.99 

2.47 

2.20 

9.29 

8.02 

WASSO.l 

0.57 

0.08 

0.22 

0.26 

0.04 

0.08 

0.05 

0.17 

0.18 

WASSOIO 

5.51 

0.81 

2.43 

2.94 

0.47 

0.84 

0.50 

1.67 

1.90 

HALAO.5 

13.47 

3.03 

7.88 

16.84 

0.81 

2.89 

0.84 

4.09 

6.23 

HALA2.0 

13.13 

3.11 

2.76 

4.54 

1.00 

1.20 

0.28 

3.62 

3.70 

SONM0.2 

22.22 

3.25 

22.71 

15.43 

19.75 

72.50 

16.26 

17.26 

23.6 

SONM2.0 

12.47 

3.66 

32.01 

42.19 

13.49 

19.84 

7.96 

8.96 

17.5 

NITRO.5 

0.12 

0.11 

2.51 

10.39 

0.35 

0.17 

0.06 

0.07 

1.72 

NTTR2.0 

0.20 

0.17 

4.09 

17.36 

0.55 

0.29 

0.11 

0.12 

2.86 

DRONO.5 

3.19 

0.90 

6.03 

6.27 

10.87 

6.46 

2.44 

3.13 

4.91 

DRON2.0 

4.47 

1.60 

11.52 

15.81 

17.05 

9.31 

3.58 

4.42 

8.47 

DRANO.5 

1.46 

0.40 

1.74 

1.65 

0.71 

2.94 

1.12 

1.31 

1.42 

DRAN2.0 

2.32 

0.68 

2.96 

3.01 

1.12 

4.58 

1.74 

1.99 

2.30 

SOPM0.2 

2.24 

0.29 

3.42 

7.15 

0.29 

10.14 

2.96 

1.07 

3.44 

SOPM2.0 

1.68 

0.21 

2.68 

4.21 

0.23 

9.15 

2.68 

0.91 

2.72 

DROPO.5 

0.70 

0.08 

1.07 

1.94 

0.09 

2.61 

2.73 

0.37 

1.20 

DROP2.0 

0.55 

0.07 

0.83 

1.61 

0.07 

2.13 

2.21 

0.29 

0.97 

DRIPO.5 

1.99 

0.27 

3.16 

6.42 

0.27 

8.43 

3.32 

0.94 

3.10 

DRIP2.0 

2.34 

0.31 

3.94 

5.74 

0.34 

14.75 

5.73 

1.28 

4.30 

SODM2.0 

0.01 

13.06 

0.24 

1.72 

0.03 

0.02 

0.01 

0.58 

1.96 

AKDDO.5 

0.00 

3.26 

0.05 

0.26 

0.01 

0.00 

0.00 

6.14 

1.21 

SUM 

5.96 

2.09 

9.82 

20.49 

2.86 

8.83 

2.84 

3.51 

7.05 

185 


Table  6.5    The  water  quality  model  coefficients  used  for  the  Charlotte  Harbor  simulation 

Coefficient 

Description 

Units 

Literature 
Range 

Charlotte 
Harbor 

(6AD)T-20 

temperature  coefficient  for 
NH4  desorption 

- 

1.08 

1.08 

(6JT-20 

temperature  coefficient  for 
algae  growth 

- 

1.01-1.2 

1.08 

(0A,)T-20 

temperature  coefficient  for 
ammonium  instability 

- 

1.08 

1.08 

(eBOD)T-20 

temperature  coefficient  for 
CBOD  oxidation 

- 

1.02-1.15 

1.08 

(6Dn)t-20 

temperature  coefficient  for 
denitrification 

- 

1.02-1.09 

1.08 

(6NN)T-20 

temperature  coefficient  for 
nitrification 

- 

1.02-1.08 

1.04 

Ood)7"20 

temperature  coefficient  for 
SON  desorption 

- 

1.08 

1.045 

/o            xT-20 
VuONM/ 

temperature  coefficient  for 
mineralization 

- 

1.02-1.09 

1.08 

re     V-20 

\VreSPS 

temperature  coefficient  for 
algae  respiration 

- 

1.045 

1.08 

(0z)T-2O 

temperature  coefficient  for 
zooplankton  growth 

- 

1.01-1.2 

1.04 

VH'a/max 

algae  maximum  growth 
rate 

1/day 

0.2-8. 

2.0-2.1 

VH'z/max 

zooplankton  maximum 
growth  rate 

1/day 

0.15-0.5 

0.16 

(NH3)air 

ammonia  concentration  in 
the  air 

Hg/L 

0.1 

0.1 

achla 

algal  carbon-chlorophyll-a 
ratio 

mg  C/mg 
Chla 

10-  112 

100 

anc 

algal  nitrogen-carbon  ratio 

mg  N  /mg  C 

0.05-0.43 

0.16 

^c 

algal  phosphorous-carbon 
ratio 

mg  N  /mg  C 

0.005-0.03 

0.025 

aoc 

algal  oxygen-carbon  ratio 

mg  02  /mg  C 

2.67 

2.667 

186 


Coefficient 

Description 

Units 

Literature 

Range 

Charlotte 
Harbor 

dan 

desorption  rate  of  adsorbed 
ammonium  nitrogen 

1/day 

- 

0.03 

don 

desorption  rate  of  adsorbed 
organic  nitrogen 

1/day 

- 

0.03 

dip 

desorption  rate  of  adsorbed 
inorganic  phosphorous 

1/day 

- 

0.02 

dop 

desorption  rate  of  adsorbed 
organic  phosphorous 

1/day 

- 

0.02 

d^l 

molecular  diffusion 
coefficient  for  dissolved 
species 

cm2/s 

4.E-6-1.E-5 

l.E-5 

Hbod 

half-saturation  constant  for 
CBOD  oxidation 

mgO-, 

0.02-5.6 

0.5 

Hn 

half-saturation  constant  for 
algae  uptake  nitrogen 

g/L 

1.5-400 

25 

Hp 

half-saturation  constant  for 
algae  uptake  phosphorous 

g/L 

1.-105 

2 

Hnit 

half-saturation  constant  for 
nitrification 

mg02 

0.1-2.0 

2.0 

hv 

Henry's  constant 

mg/L-atm 

43.8 

45 

Is 

optimum  light  intensity  for 
algal  growth 

|J.E  /m2  /s 

300-350 

350 

Kax 

excretion  rate  by  algae 

1/day 

0.02-0.24 

0.06 

Kas 

mortality  rate  of  algae 

1/day 

0.2-0.22 

0.07 

KAI 

ammonia  conversion  rate 
constant 

1/day 

0.01-0.1 

0.01-0.02 

KD 

CBOD  oxidation  rate 

1/day 

0.02-0.6 

0.05 

^DN 

denitrification  rate  constant 

1/day 

0.02-1.0 

0.09 

^NN 

nitrification  rate  constant 

1/day 

0.004-0.11 

0.08 

Kon 

rate  of  ammonification  of 
SON 

1/day 

0.001-1.0 

0.015 

187 


Coefficient 

Description 

Units 

Literature 
Range 

Charlotte 
Harbor 

Kop 

rate  of  mineralization  of 
SOP 

1/day 

0.001-0.6 

0.02 

KVoi 

rate  constant  for  nitrogen 
volatilization 

1/day 

3.5-9.0 

7. 

Kzs 

mortality  rate  of 
zooplankton 

1/day 

0.001-0.36 

0.02 

Kzx 

excretion  rate  of 
zooplankton 

1/day 

0.003-0.075 

0.01 

Pan 

partition  coefficient 
between  SAN  and  PEN 

I/Jig 

.5E-7-1.E-5 

1E-4 

Pon 

partition  coefficient 
between  SON  and  PON 

1/M-S 

1.0E-5 

LE-5 

Pip 

partition  coefficient 
between  SRP  and  PEP 

i/ng 

- 

1E-5 

Pop 

partition  coefficient  of 
SOP  and  POP 

m 

- 

E-4 

W^CBOD 

CBOD  settling  velocity 

cm/s 

- 

0.01 

^*^algae 

algae  settling  velocity 

cm/s 

0.0-300. 

10. 

Figure  6.8  contains  the  scattering  plots  for  the  calibration  period.  The  location  of 
circles  indicates  the  correlation  between  model  predictions  and  observed  data.  A  perfect 
match  between  model  and  observed  data  is  indicated  by  the  diagonal  line  on  each  graph.  The 
circle  above  the  line  is  over  predicting  the  observation.  Circles  below  the  line  indicate  that 
the  model  is  under  predicting  the  observation.  Shown  with  each  plot  are  the  root  mean 
square  (RMS)  error  and  correlation  coefficient  (R2)  mentioned  before.  Values  shown  in 
parenthesis  are  RMS  errors  normalized  with  maximum  measured  data. 

Overall,  the  model  results  show  reasonable  RMS  errors  for  all  constituents,  although 
the  R2  values  are  not  very  high.  The  normalized  RMS  errors  are  less  than  50%,  except  NH4, 
P04,  and  TSS.  Dissolved  oxygen  has  the  best  agreement  with  measured  data  as  shown  by 


188 
the  RMS  error.  Total  suspended  sediment  concentration  changed  quickly  with  time  on  the 
order  of  minutes  and  hours,  hence  comparison  with  monthly  data  did  not  show  reasonable 
RMS  error.  Dissolved  ammonium  nitrogen  and  soluble  reactive  phosphorous  are  strongly 
affected  by  sorption  and  desorption  processes,  while  particulate  species  are  affected  by 
dynamic  processes  of  settling  and  erosion  of  suspended  sediment.  The  water  quality  model 
calculates  CBOD  instead  of  total  organic  carbon.  The  simulated  CBOD  was  compared  with 
measured  total  organic  carbon  because  there  is  no  measured  CBOD  data.  Hence,  the  amount 
of  CBOD  was  under  predicted  as  shown  in  the  scatter  plot,  due  to  the  difference  between 
CBOD  and  total  organic  carbon  concentration  in  the  system. 

The  correlation  coefficient  (R2)  measures  the  strength  of  linear  association  between 
simulated  and  measured  data.  Correlation  coefficients  for  all  species  vary  form  0.2  to  0.88. 


189 


A)  Chrolophyll  a 

25 


B)  Dissolved  Oxygen  (DO) 


RMS=7.51  (33.56%) 
R2     =0.519 


20 

£      15 
w 

1     10 


C)  Total  Kjeldahl  Nitrogen  (TKN) 


500  1000 

measured 


IU 

RMS=  1.40  (17.63%)         #X 

R2     =0.308           **       ,X 

*  •<*<%/. 

•*  ,4r 

•  ui?v. 

(1) 

3 

5 

E 

•    -X'  • 

c/) 

n 

5         10        15        20        25 

measured 


IOUU 

RMS=0.228(21.98 

%)         / 

R2     =0.462 

•^r 

• 

simulated 

o                  o 

0                    o                    o 

.   .  *x • 

*3i»  •  •  • 

• 

• 
• 

• 
i     ,     ,     ,     , 

measured 

D)  Dissolved  Ammonium  Nitrogen  (NH ,) 

250 


10 


RMS=  0.040  (44.48%) 
R2     =  0.362 


150( 


100  200 

measured 


Figure  6.8  The  scatter  plots  for  water  quality  constituents  during  calibration  period 


190 


E)  Total  Phosphorous 

450 


400 
350 


RMS=  0.0783  (31.26%) 
7  R2     =0.242 


F)  Soluble  Reactive  Phosphorous  (P04) 

250 1 


100       200       300       400 

measured 

G)  Total  Organic  Carbon  (CBOD) 


1  ■•*    »  i    i    ■    I    ■    .    i    i    1    i    i    i    ■    I    i    ■ 


£U 

RMS=  3.320  (32.13%)          / 

R2     =  0.806                       »A* 

15 

/>•'• 

■a 

f                                • 

0) 

/                    *           • 

*•> 

X           • 

m 

/    •  .  • 

=j 

10 

s           • 

£ 

/     •*:.# 

w 

A*  •  \ . 

5 
n 

/ J*  *'»  * 
AW. 

0         50       100      150      200      250 

measured 

H)  Total  Suspended  Sediment  (TSS) 

150 


RMS=  25.67  (36.45%) 
R2     =0.124 


"0  5  10  15 

measured 

Figure  6.8  Continued 


20 


««h< 


50  100 

measured 


150 


191 
6.3.2  Results  of  1996  Water  Quality  Simulation 

The  temporal  variation  of  water  quality  species  at  each  measured  station  is  compared 
with  measured  data  in  Figures  6.9  to  6.20.  The  thicker  solid  lines  represent  the  simulated 
water  quality  parameters  near  surface  and  dashed  lines  are  those  for  bottom  layer.  The 
measured  water  quality  parameters  are  represented  with  rectangular  symbols  which  represent 
several  vertical  layers.  The  numbers  of  vertical  layer  of  measured  data  are  all  different  for 
each  species  and  each  station.  The  measured  data  were  plotted  at  all  available  vertical 
measured  location.  The  results  show  simulated  and  measured  Chlorophyll_a,  dissolved 
oxygen,  TKN,  dissolved  ammonium  nitrogen,  total  phosphorous,  and  soluble  reactive 
phosphorous.  During  the  simulation  period,  there  was  relatively  little  temporal  variation  in 
the  water  quality  parameters.  One  major  exception  is  at  the  CH-005  and  CH-006  station 
which  showed  a  significant  vertical  stratification  of  dissolved  oxygen  concentration. 

The  dissolved  oxygen  processes  are  important  in  any  aquatic  environment,  because 
living  organisms  depend  on  oxygen  in  one  form  or  another  to  maintain  their  metabolic 
processes.  In  Charlotte  Harbor  estuarine  system,  bottom  water  hypoxia  has  been  reported 
periodically  by  Environmental  Quality  Laboratory  (EQL)  since  the  mid-1970  (Heyl,  1996). 
In  the  water  quality  model,  dissolved  oxygen  is  a  function  of  photosynthesis  and  respiration 
by  phytoplankton  organisms,  sediment  oxygen  demand,  reaeration,  nitrification  and 
denitrification,  decomposition  of  organic  matter,  tide  and  wind  mixing,  and  river  loading.  As 
shown  in  Figures  6.9  to  6.20,  concentrations  of  dissolved  oxygen  exhibit  a  temporal  and 
vertical  variation  in  response  to  variations  in  phytoplankton  biomass  and  nitrogen  species  in 
the  northern  part  of  Charlotte  Harbor,  while  relatively  little  variation  in  the  southern  part  of 
Charlotte  Harbor  during  the  simulation  period.  Figure  6-21  shows  the  snapshots  of  the  near 


192 
simulated  bottom  dissolved  oxygen  distribution  in  the  study  area  on  August  22, 1996  (at  the 
end  of  90-day  simulation).  The  result  corresponds  well  with  the  measured  low  (<  2mg/L) 
dissolved  oxygen  in  the  bottom  water  at  the  Peace  River  mouth.  The  rest  of  estuary  does  not 
show  low  dissolved  oxygen  less  than  2  mg/1. 

Due  to  wind  and  tidal  mixing  and  the  shallow  water  depth,  the  Charlotte  Harbor 
estuarine  system  generally  exhibits  a  vertically  well  mixed  distribution  of  DO.  In  the  upper 
Charlotte  Harbor,  where  the  lowest  DO  in  the  system  is  usually  found,  stratification  may 
occur  and  a  pycnocline  may  form  if  the  condition  is  right  -  high  river  inflow  from  the  Peace 
river  and  low  wind  mixing  (Sheng  and  Park,  2002).  With  this  strong  stratification,  surface 
water  DO  from  reaeration  would  be  blocked  and  high  sediment  oxygen  demand  in  summer 
season  could  create  the  hypoxia  observed  in  this  area.  Model  results  showed  that  strong 
stratification  and  low  DO  developed  during  the  simulation  period. 

Phytoplankton  dynamics  are  very  important  processes  simulated  in  this  study.  It  is 
closely  related  to  the  nutrient  recycling  through  uptake  during  growth,  and  excretion/decay 
during  respiration.  The  measured  phytoplankton  as  algal  mass  per  volume,  was  converted 
to  phytoplankton  carbon  with  an  algae  to  carbon  ratio  of  100.  The  chlorophyll_a 
concentrations  of  Myakka  and  Peace  River  have  maximum  value  of  on  July  23  (Julian  Day 
204)  and  those  for  the  other  stations  have  no  trends  during  the  simulation  periods.  The 
highest  daily  fluctuation  of  chlorophyll_a  is  occurred  in  HB006  due  to  strong  tidal 
fluctuation  and  high  chlorophyll_a  concentration  from  the  Caloosahatchee  River. 

In  order  to  compare  the  nitrogen  cycle  simulations  with  the  nitrogen  species  data 
provided  by  EPA,  simulated  soluble  organic  nitrogen,  dissolved  ammonium  and  ammonia 
nitrogen  concentrations  are  combined  and  compared  with  the  EPA  total  Kjeld  nitrogen  data. 


193 
The  dissolved  ammonium  nitrogen  data  were  also  compared  with  measured  data.  For 
phosphorous  species,  total  phosphorous  concentration  and  soluble  reactive  phosphorous  are 
used  to  compare  with  the  EPA  data.  According  to  time  series  plot  of  each  species  in  Figures 
6.9  to  6.20,  model  results  appear  to  capture  the  overall  trend  of  the  EPA  data. 

Figure  6-22  shows  the  snapshots  of  the  near  surface  chlorophyll_a  distribution  in 
study  area  for  August  22  (the  end  of  90-day  simulation).  Only  the  Peace  River  mouth  area 
exhibits  very  high  chlorophyll_a  concentration  of  3000p-g  (phytoplankton  carbon).  Higher 
phytoplankton  concentrations  are  generally  found  near  the  river  mouths  and  low 
concentrations  are  found  in  the  Gulf  of  Mexico. 

Figures  6.23  to  6.26  show  the  snapshots  of  the  near  surface  dissolved  ammonium 
nitrogen,  soluble  organic  nitrogen,  soluble  reactive  phosphorous,  and  soluble  organic 
phosphorous  distributions  in  the  study  area  at  2  pm  on  August  21, 1996  (at  the  end  of  90-day 
simulation).  It  is  interesting  to  note  that  high  dissolved  ammonia  concentrations  but  low 
dissolved  organic  nitrogen  concentrations  are  found  in  the  area  between  upper  Charlotte 
Harbor  and  the  Boca  Grande  Pass.  High  concentrations  of  particulate  organic  nitrogen  and 
adsorbed  ammonium  are  first  resuspended  by  the  strong  currents  in  the  Boca  Grande  Pass 
area,  causing  high  dissolved  ammonium  concentration  in  the  water  column  which  are  then 
transported  towards  the  upper  Charlotte  Harbor  area.  High  concentrations  of  phosphorous 
species  are  found  in  the  upper  Charlotte  Harbor  because  Peace  River  drains  the  Hawthorn 
phosphatic  formations. 


194 


160 


180 


200 


220 


|   30 
<o   20 


Q. 
O 


o 


10 


160 


180 


200 


220 


2000 


§  1500 

— '  1000  - 

H  500  • 


160 


i^w^wv»/^rf' ^^m^^^vvv-J~^~-"''^v^.'Vvwm^ 


180 


200 


220 


500 
g.  400 
J  300 
x-  200 
Z  100 
0 


160  180  200  "         220       "~ 


1000 

800 

■a 

ROO 

d 

Q. 

400 

t- 

200 

0 

500 

^^ 

400 

"5> 

-i 

300 

** 

o 

200 

□. 

100 

160 


180 


200 


220 


160  180  200  220 

Figure  6.9  Temporal  water  quality  variations  at  CH002  station  in  1996. 


195 


10  r 


O 

Q    2 

0 


*rwww\^I^ 


»wv  \\jf  'vwq 


160 


180 


200 


220 


160 


180 


200 


220 


2000 


f*  1500 


500 

0 


1000  *^^-^' 


/WV\*|^ 


500  r 
.-.  400 
J  300 
£•  200 
z  100 
0 


160 


180 


200 


220 


I 

uMMjU 

■ 

a(L « 

mMmaw 

iMMMy^M^yvvWito 

160 

180 

200 

■                       220                  " 

1000 

800 

"9>  600 

i   400 

200 

0 


500 


160 


180 


200 


220 


160  180  200  220 

Figure  6.10  Temporal  water  quality  variations  at  CH004  station  in  1996. 


196 


\m 


■• 


180  200 


y^mfki'V)0 


220 


160 


"&> 

au 

-i 

"■ — 

■ 

oj 

20 

>. 

.c 

CL 

o 

c 

o 

10 

JP«MA/ 

tfjtW/Wwwi 

■ 

™rv«*\\A/v'w^ 

WJMK 

mMJ}M^ 

flL  .iA-' 

.c 

0 

u 

160 

180 

200 

220 

2000 


f*  150° 
-'  1000 


1^^-^.^WaV 


P      500 

0 

500 
^     400 
J     300  f- 
3-'    200 
Z      100 


160 


180 


200 


"220" 


0  m»AA*AAWjmmy*^^ 


200 


220 


160 


180 


200 


220 


500 


160  180  200  220 

Figure  6.1 1  Temporal  water  quality  variations  at  CH005  station  in  1996. 


197 


160 


180 


200 


220 


re 

>> 
.c 
D. 

2 
o 

.c 
o 


JU 

20 

■ 

10 

0 

^!«VAAflMW!©5M^ 

1    " 

,vJkNW3v«JlwY 

vaII  M/       -  -^^    ^ArvvJ 

T\-V?  **\- 

^-H"— vvK\j 

L^-''r 

160 

180 

200 

220 

2000 


1*  1500 
*  1000 
£  500^ 


ltfiMwJW^rv^'tJ*JWM^^ 


500 


160 


160 


180 


200 


220 


1UUU 

800 

3 

a 

i- 

600 
400 
200 

0 

.aft,     .         miw/ 

ICuhJU                    ■  ftt 

w^^^j^^^A^ 

160 

180 

200 

220 

500  r 


160  180  200  220 

Figure  6.12  Temporal  water  quality  variations  at  CH006  station  in  1996. 


198 


E 

O 
Q 


10 
8 
6 
4 
2 
0 


500 
g.  400 
1  300 


X 

200 

z 

100 

0 

1000 

f 

2 
a. 

800 
600 
400 
200 

0 

500 

1 

400 
300 

O 
a 

200 
100 

0 

T"'* 


160 


180 


200 


220 


160 


160 180 200  220 


160 


180 


200 


220 


160  180  200  220 

Figure  6.13  Temporal  water  quality  variations  at  CH007  station  in  1996. 


199 


O 
a 


|  30 

m  20  ■ 

> 

|  10 

o 

5     o 


160 


180 


200 


220 


■g*^^^^^ 


160 


180 


200 


220 


2000 


U  1500 

■3"  1000 

z 

H      500 


500 
^  400 
t  300 
x"»    200 

Z      100 


KJLS**-*** 


^^/V^^'-J 


160 


180 


200 


220 


laaaaaftttMawwvflt^^ 


160 


180 


200 


220 


1000 

800 

■& 

600 

3 

Q- 

400 

1- 

200 

0 

500 

/— s 

400 

-1 

300 

o 

200 

a 

100 

0 

160 


180 


200 


220 


160  180  200  220 

Figure  6.14  Temporal  water  quality  variations  at  CH09B  station  in  1996. 


200 


10 

«-.    8 


g    6  fc^v^^^5^^^ 


O 
Q 


5 

>. 

a. 
o 


4  t 
2 


160 


180 


200 


220 


160 


180 


200 


220 


2000  r 


ffc  1500 
w  1000 


500 


BjVA^V,— j~*^V*WW*. 


NVM^WW^f 


r^JvJWV-v^v/'/*/v^W' 


,^.^^Vv*WVW«^'ww<M^wn^/^'^^*^^^ 


160 


180" 


200 


220 


500  F 
j-.  400 
J  300  f 
_.*  200 


°"^      ~T60  180     '       200  220 


1000 

800 

■ft 

600 

Ti 

a. 

400 

i- 

200 

0 

500 

^^ 

400 

a 

300 

O 

200 

Q. 

100 

0 

160 


180 


200 


220 


160  180  200  220 

Figure  6.15  Temporal  water  quality  variations  at  CH009  station  in  1996. 


201 


■a 

E 
o 

D 


10 

8 

6  | 
4 

2 
0 


160 


180 


200 


220 


5    30 


TO 
~>. 

SI 

a. 

2 
o 

.c 
O 


20 


10 


■a. 


5 

I 

z 


0 

2000 

1500 

1000 

500 

0 

500 
400 
300 
200 
100 


T60-         180  200  220 


***~~*<w*^*i»*>iW<t1fiM^^ 


160 


180 


200 


220 


™^  1  «n      '         Tori  '  oAa  ™   '         oon         " 


160 


180 


200 


220 


=5. 


1000 
800 
600 


o 

a. 


400 
200 

0 

500 
400 
300 
200 


|  l^-^WW^'^W 


160 


180 


200 


220 


l|~~""^^^ 


Figure 


160  180  200  220 

6.15  Temporal  water  quality  variations  at  CH010  station  in  1996. 


202 


10 

8 

E 

6 

O 

4 

Q 

2 

0 

I 

30 

s 

"— ^ 

ro 

20 

>. 

.n 

Q. 
O 

10 

O 

Z 

o 

0 

2000 

0 

1500 

«.        a     hi 

At    iLllf 

mi  1 1 AA  a 

^VvvH^ 

*%P 

^jW^ 

^#% 

m!0&i^d>MmN 

■ 

160 

180 

200 

220 

160 


180 


200 


220 


i 

—  1000 

z 

H   500 


'VVVwA^^^ 


160 


180 


200 


220 


500  r 


1 

3 

400 
300 

z 

200 
100 

0 

1000 

f 

Q. 

H 

800 
600 
400 
200 

0 

500 

1 

2 

400 
300 

if 

O 

Q. 

200 
100 

0 

/^M/I/VI/ma/mm/m^^ 


160 


180 


200 


220 


160 


180 


200 


220 


160  180  200 

Figure  6.17  Temporal  water  quality  variations  at  HB002  station  in  1996 


220 


203 


10 

Q 

o   4 


160 


|   30 

n   20 

>. 

f   10 

5  o 

2000 


fi  1500 
&   1000 

z 

K   500  [■ 


160 


kW.'.^*™^ 


0 

500 
-.  400 
1  300 
j?  200  ■ 
Z   100  • 
0 


160 


160 


1000 
^  800 
"^  600 
a   400 

H   200  t 


500 
^  400 
§  300 
Q'  200 

a   100 

0 


160 


180 


200 


220 


180 


200 


220 


T80~  200 


WW 


220 


180 


200 


220 


180 


200 


220 


'^W-Vw'Bs,~*^^ 


160  180  200  220 

Figure  6.18  Temporal  water  quality  variations  at  HB006  station  in  1996. 


204 


E 

O 
Q 


10 
8 
6 
4 
2 
0 


f**ZXf^ 


^$^*^ 


K/\M 


ns/vj 


160 


180 


"200" 


220 


0  160  ~~^80~~  ~200  220" 

Figure  6.19  Temporal  water  quality  variations  at  HB007  station  in  1996. 


205 


"a 
J, 

o 


160 


180 


200 


220 


■a 


Q. 
O 

I— 

O 

r 
o 


30 

20 

10 

0 

2000 


=•  1500 

-"  1000 


500 

0 

500 
400 
300 
200 
100 
0 

1000 
800 
600 
400 
200  ■ 
0 

500 
400 
300 
200 
100 
0 


160 


180 


200 


220 


■a. 


a. 


160 


■a 

3 

t 

o 
a. 


160  180  200    ~  220 

Figure  6.20  Temporal  water  quality  variations  at  CH013  station  in  1996. 


206 


Near-Surface  Dissolved  Oxygen  (mg/L) 


Near-Bottom  Dissolved  Oxygen  (mg/L) 


August  21, 1996 


5.  \ 

Figure  6.21  Simulated  dissolved  oxygen  concentration  in  Charlotte  Harbor  estuarine 
system  on  August  21,  1996. 


207 


Near-Surface  Chlorophyll  a  (ug/L) 


Near-Bottom  Chlorophyll  a  (ug/L) 


Figure  6.22  Simulated  Chlorophyll_a  concentration  in  Charlotte  Harbor  estuarine  system 
on  August  21,  1996. 


208 


Near-Surface  NH4  (ug/L) 


Near-Bottom  NH4  (ug/L) 


Figure  6.23  Simulated  dissolved  ammonium  nitrogen  concentration  in  Charlotte  Harbor 
estuarine  system  on  August  21,  1996. 


209 


Near-Surface  SON  (ng/L) 


Near-Bottom  SON  (ug/L) 


Figure  6.24  Simulated  soluble  organic  nitrogen  concentration  in  Charlotte  Harbor 
estuarine  system  on  August  21,  1996. 


210 


Near-Surface  SRP  (ng/L) 


Near-Bottom  SRP  (|ig/L) 


500        / 
400        \ 


\ 


X  \ 


J 


500 
400 
300 
200 
100 


V 


August  21, 1996 


Figure  6.25  Simulated  soluble  reactive  phosphorous  concentration  in  Charlotte 
Harbor  estuarine  system  on  August  21,  1996. 


211 


Near-Surface  SOP  (ng/L) 


Near-Bottom  SOP  (ug/L) 


60 
50 
40 
30 
20 
10 


August  21, 1996 


v  *^ 


Figure  6.26  Simulated  soluble  organic  phosphorous  concentration  in  Charlotte 
Harbor  estuarine  system  on  August  21,  1996. 


212 
6.4  Simulations  of  Water  Quality  in  2000 

Model  simulation  of  circulation,  sediment  transport  and  water  quality  dynamics  in 
Charlotte  Harbor  estuarine  system  during  the  summer  1996  was  conducted  as  calibration 
process  tunning  of  model  parameters  and  inputs  followed  systematic  calibration  procedure. 
Based  on  the  results  presented,  the  water  quality  model  was  considered  calibrated  and 
validated,  and  additional  one  year  simulation  was  performed  to  validate  the  model.  This 
validated  model  was  used  to  assess  the  effects  of  Sanibel  Causeway  and  the  navigation 
channel  in  San  Carlos  Bay  as  well  as  river  load  reductions. 
6.4.1  Validation 

For  validation  runs,  hydrodynamic  and  water  quality  model  coefficients  were  held 
fixed  at  the  calibration  values,  and  results  of  a  one-year  2000  simulation  were  compared  with 
field  data  collected  in  January  to  December  2000.  The  data  include  phytoplankton,  nitrogen, 
phosphorous,  dissolved  oxygen  at  several  locations  in  the  Charlotte  Harbor  estuarine  system. 

Figure  6.27  contains  calibration  period  scattering  plots  which  include  the  root  mean 
square  (RMS)  error  and  correlation  coefficient  (R2).  Values  shown  in  parenthesis  are 
normalized  RMS  errors  with  maximum  measured  data.  Without  adjusting  any  water  quality 
model  coefficients,  the  normalized  RMS  errors  are  less  than  45%.  Dissolved  oxygen  has  the 
best  agreement  with  measured  data  as  shown  by  the  RMS  error.  Correlation  coefficients  for 
all  species  vary  form  0. 124  to  0.80.  These  plots  allow  us  to  assess  weaknesses  of  the  model, 
and  to  suggest  areas  needing  further  improvement. 


213 


A)  Chrolophyll  a 

25 


to 


RMS=  7.51  (33.56  %) 
R2     =0.519 

20h   * 

■a 

$>      15- 
co 

E-'  I-  / 

-  iot.  l-jyzs**  •  •  • 

• . vvn>s  •   • • 





"0  5         10        15        20 

measured 

C)  Total  Kjeldahl  Nitrogen  (TKN) 

1500 


■D1000 

3 

E 

«   500 


B)  Dissolved  Oxygen  (DO) 

10 


CD 

£ 

CO 


RMS=  1.40  (17.63%) 
R2     =  0.308 


25 


0, 


RMS=0.228(21.98 

%)         / 

R2 

=  0.462 

mf 

• 

•  • 

• 
• 

•  I  •    •  4  Sm  • 

•  *y * 

Vm 

9 

•; 

• 
• 

• 

" 

pA- 

r        * 

/ 

,,!,,,, 

,     ,     ,     , 

measured 

D)  Dissolved  Ammonium  Nitrogen  (NH4) 

250 


RMS=  0.040  (44.48  %) 
R2     =  0.362 


500  1000  150(  0  100  200 

measured  measured 

Figure  6.27  The  scatter  plots  for  water  quality  constituents  in  2000. 


214 


E)  Total  Phosphorous 

450 


400 
350 


RMS=  0.0783  (31.26%) 
R2     =  0.242 


■o 
E 


RMS=  3.320  (32.13%) 
R2     =  0.806 


100       200       300       400 

measured 

G)  Total  Organic  Carbon  (CBOD) 

20 


5  10  15 

measured 

Figure  6.27  continued 


20 


F)  Soluble  Reactive  Phosphorous  (P04) 

250 


RMS=  0.0294  (42.87%) 
R2     =  0.280 


200- 


50       100      150     200     250 

measured 

H)  Total  Suspended  Sediment  (TSS) 

150 


RMS=  25.67  (36.45  %) 
R2     =0.124 


50  100 

measured 


150 


215 
6.4.2  Results  of  2000  Water  Quality  Simulation 

The  simulated  water  quality  species  at  each  measured  station  in  2000  as  shown  with 
measured  data  in  Figure  6.28  to  Figure  6.39.  During  the  simulation  period,  the  water  quality 
data  show  the  seasonal  variation,  and  these  seasonal  variations  are  produced  quite  well  by 
the  water  quality  model.  In  the  water  quality  model,  dissolved  oxygen  is  a  function  of 
photosynthesis  and  respiration  by  phytoplankton  organisms,  sediment  oxygen  demand, 
reaeration,  nitrification  and  denitrification,  decomposition  of  organic  matter,  tide  and  wind 
mixing,  and  river  loading.  As  shown  in  Figures  6.28  to  6.39,  simulated  dissolved  oxygen 
concentrations  show  spatial  and  temporal  variations  in  reasonable  agreement  with  measured 
data.  The  average  monthly  near-surface  concentrations  declined  from  8.5  to  6.7  mg/L  from 
January  to  July  and  then  began  to  rise  in  upper  Charlotte  Harbor.  Near-bottom  average 
monthly  concentrations  in  this  area  were  highest  in  February,  declined  slowly  through  May, 
and  then  declined  more  rapidly  until  September.  The  hypoxia  conditions  during  summer  are 
attributed  to  strong  stratification,  which  cause  restricted  re-aeration,  and  SOD.  After  the 
breakup  of  the  stratification,  the  DO  concentration  increased  from  October  to  December. 

Algae  photosynthesis  and  re-aeration  maintain  surface  dissolved  oxygen  level,  while 
vertical  mixing  controls  the  transfer  of  dissolved  oxygen  to  bottom  water.  Sediment  oxygen 
demand  is  a  function  of  temperature  which  has  strong  seasonal  variation.  The  high  water 
temperature  in  summer  season  will  increase  SOD  in  bottom  sediment.  If  there  is  no 
stratification,  surface  water  dissolved  oxygen  will  become  mixed  with  bottom  water 
dissolved  oxygen  quickly.  The  Charlotte  Harbor  estuarine  system  generally  exhibits  a 
vertically  well  mixed  distribution  of  DO  due  to  wind  and  tidal  mixing  and  the  shallow  water 
depth.      The  dissolved  oxygen  in  most  part  of  the  estuary  does  not  show  any  strong 


216 
stratification  except  in  Peace  and  Caloosahatchee  Rivers.  In  the  upper  Charlotte  Harbor, 
which  usually  has  the  lowest  level  of  DO,  some  stratification  may  occur  due  to  high 
consumption  by  SOD  near  the  bottom  and  super-saturation  near  the  surface  with  strong  river 
discharge  from  Peace  River.  To  quantify  the  causes  of  the  bottom  water  hypoxia  in  this  area, 
dissolved  oxygen  concentration  at  CH006  was  compared  with  river  discharge  at  Peace  River, 
salinity,  temperature  and,  reaeration  and  SOD  fluxes  in  Figure  6.40.  Although,  the  SOD  flux 
is  higher  than  the  reaeration  flux  in  summer  (Julian  Day  120  to  280)  due  to  high 
temperature,  the  DO  stratification  did  not  occur  during  this  period.  The  DO  stratification 
period  (Julian  Day  220-280)  matches  with  the  salinity  stratification  caused  by  strong  river 
discharge  from  Peace  River. 

To  compare  salinity  stratification  and  vertical  DO  distribution,  Simulated 
longitudinal-vertical  salinity  and  dissolved  oxygen  concentration  along  the  Peace  River  at 
1  pm  on  June  18  (Julian  Day  170)  and  October  6  (Julian  Day  280),  2000  were  plotted  in 
Figures  6.41  and  6.42,  respectively.  The  salinity  at  these  two  time  period  represent  the  effect 
of  salinity  stratification  on  vertical  DO  distribution  since  the  SOD  and  re-aeration  fluxes  are 
similar  at  these  periods.  The  results  show  the  strong  relationship  between  salinity  and  DO 
stratifications.  Therefore,  the  hypoxia  in  the  upper  Charlotte  Harbor  is  primarily  caused  by 
the  combination  effects  of  SOD  and  stratification  with  strong  river  discharge  and  high 
temperature. 

In  the  Caloosahatchee  River  upstream  near  the  S79  (CES02),  there  is  very  strong 
daily  fluctuation  in  the  surface  dissolved  oxygen  concentration  due  to  tidal  fluctuation  and 
high  dissolved  oxygen  concentration  from  the  Caloosahatchee  River.  Chlorophyll_a 
concentrations  ranged  from  1  to  100  mg/L  and  averaged  8.5  mg/L.  Both  productivity  and 


217 
biomass  were  greater  during  summer  near  the  mouth  of  tidal  rivers  which  has  middle  range 

salinity  of  6  to  12  ppt.  The  chlorophyll_a  concentration  of  Caloosahatchee  River  has  a 
maximum  value  of  98  mg/L  on  July  12  (Julian  Day  193),  2000  and  then  quickly  drop  to  5 
mg/L  because  of  low  chlorophyll_a  concentration  of  river  loading  and  fast  flushing  due  to 
strong  river  discharge  from  Caloosahatchee  River.  Simulated  nitrogen  and  phosphorous 
concentrations  appear  to  capture  the  overall  trend  of  the  measured  data  collected  by 
SFWMD  and  SWFWMD. 

Figures  6.43  and  6.44  show  the  snapshots  of  the  near  surface  chlorophyll_a  and  the 
near  bottom  dissolved  oxygen  concentration  distribution  in  the  study  area  on  February  9, 
May  9,  August  7,  and  November  5,  to  represent  seasonal  characteristics.  In  February, 
phytoplankton  was  low  in  the  entire  estuary  and  DO  was  a  generally  high.  In  May, 
phytoplankton  significantly  increased  in  the  Caloosahatchee  River  due  to  strong  river 
discharge  with  high  nutrient  concentrations.  Hence,  DO  was  reduced  substantially  in  all 
portions  but  was  higher  in  the  north  than  in  the  south,  except  in  Pine  Island  Sound.  In 
August,  there  was  significant  increase  in  phytoplankton  and  decrease  in  dissolved  oxygen 
concentration  in  the  bottom  water.  This  low  dissolved  oxygen  in  bottom  water  reached  the 
north  portion  of  upper  Charlotte  Harbor.  The  new  water  quality  model,  which  contains 
improved  dissolved  oxygen  processes  at  the  air/sea  interface  and  water-sediment  interface, 
successfully  reproduced  the  bottom  water  hypoxia  in  both  temporal  and  spatial  plots. 


218 


1 

E 

o 

Q 


I       30 

3. 


>» 

a 
o 

S 

o 


20 

10 

0 

1500 

1000 
500 


^— i -*-«-^^^ 


^»^ 


•6) 


200 
150 
100  - 
50  ■ 

400 
300 
200 

100 

0 

200 


=;  150 

■3  100 
o 

0.  50 


a. 


**** 


100 


"500" 


300 


-^^^>-^^_ 


100 


^00 


300 


[- 

100 

200 

300 

■ 

1 — .1^ 

"*"**«» 

""■■  DlJlljlh 

1                          -y. 
in    iiiiiiiiiiiiiJiiiJ^rfP^ 

1    OKU 

«*JLgI 

"^ H^^»»  .     a.   .^._ ^B 

100 


200 


300 


^^^Jj>,y*L-, 


100  200  300 

Figure  6.28  Temporal  water  quality  variations  at  CH002  station  in  2000. 


219 


§ 

o 

Q 


|       3° 
3. 


TO 

>. 
.C 
Q. 

2 

o 

O 


100 


200 


20 

10 

0 

1500 

1000 


ksr^*^^^^^^ 


*      500 

0 
200 


i 

z 


150 
100 

50 


oPVwJl 


5 


■B) 

if 

o 


100 


200 


100 


200 


300 


300 


300 


WHM'Jftl 


100 


200 


iiiiwmwimwi 


A 


300 


Figure 


100  200  300 

6.29  Temporal  water  quality  variations  at  CH004  station  in  2000. 


220 


O 
Q 


■a 

a, 


Q. 
O 

i- 

o 

i 

o 


1 

3 


30 

20 

10 

0 

1500 

1000 

500 

0 

200 
150 

100 

I* 

Z   50 

0 

400 

O  300 

3  200 
Q. 

*-     100 


i*-ff^l — nr^J^^^*^^ 


100 


200 


300 


/-T 


»**"*" 


100 


200 


"300" 


"5> 
2 


wJLfciriffMjLt^r  riv 


100 


200 


300 


•-if"**^^ 


100 


200 


300 


5 

o 

a 


Figure 


100  200  300 

6.30  Temporal  water  quality  variations  at  CH005  station  in  2000. 


221 


■5) 
E 

O 
a 


%      30 


100 


200 


300 


Q. 

2 
o 

o 


20 


10 


Piftw**   », 


100 


200 


300 


1500 


"S>  1000 

1      500 

I- 


200 


yw' 


,_^^|^ 


«U#W/»fc 


m  i   «njl»  ,w   ^  i^^'^i  #  ■*![  in      i 


100 


200 


300 


100 


200 


300 


400  r 


~     300  [ 

% 

3     200 

Q. 

*-     100 


^t-^-»*-^^jl^a*^^ — . 


100 


200 


300 


200 


=;     150 
o) 

«     100 

o 

Q.        50 


100 


200 


300 


Figure  6.31  Temporal  water  quality  variations  at  CH006  station  in  2000. 


222 


10 

£ 

8 
6 

-^ 

o 

4 

Q 

2 

0 

1 

30 

3 

-w' 

TO 

20 

>< 

J= 

Q. 
O 

10 

i- 

O 

£ 

o 

0 

1500 

"S>  1000 
*   500 

I- 


■5) 


■a. 


"ft 

o 

Q. 


150 
100 

50 
0 

400 
300 
200 
100 

0 

200  r 

150 

100 

50 
0 


100 


200 


300 


100 


200 


300 


~_*f-'-"vJ 


100 


200 


300 


k^Nrt* 


100 


200 


300 


JUttAm "  i - 


ll^Mt^^k 


100  200  30( 

Figure  6.32  Temporal  water  quality  variations  at  CH007  station  in  2000. 


223 


100 


200 


300 


% 

30 

3 

w 

TO 

20 

>< 

.C 

Q. 
O 

10 

O 

.C 

o 

0 

1500 

^ 

"& 

1000 

d 

,— •* 

7 

* 

600 

h- 

'*>il*i«mmm*M^m>i'm^f0>^ 


«^*^^^ 


100 


200 


300 


>„^~~»v-'" 


100 


200 


300 


200 


f 

2 

I 
2 

150 
100 

50 

0 

400 

1 

zt 

300 
200 

i- 

100 

)J^^k^mm»^^ 


100 


200 


300 


100 


200 


300 


200 


^     150 


3 

O 

a 


100 

50 
0 


fr   ^""""If**^*"^ 


Figure 


100  200  300 

6.33  Temporal  water  quality  variations  at  CH008  station  in  2000. 


224 


E 

O 
Q 


10 
8 
6 
4 
2 
0 


100 


200 


300 


i   30 

•   20 

§"   10 


1500 


"3)  1000 
*   500 


^jj**<lim^m»<,'i-»iii«*^ 


riMHP 


H^.^^**--^^^ 


100 


200 


300 


0 
200 


=;  150 
*  100 


_■ ■   "***J^'II>*J"> I   '   |L  M  iii  Irt.nf-J-l 


100 


200 


"300" 


I 

z 


50 
0 

400 


TOO      '       '   200   '    '    '    '   500      ~ 


ss-  300 

3  200 

Q. 

*-  100  ■ 

0 
200 


g  150 

3  100 

o 

Q.  50 


100 


200 


300 


>^JL*^ I 


100  200  300 

Figure  6.34  Temporal  water  quality  variations  at  CH009  station  in  2000. 


225 


f 

o 

Q 


"ft 

(0 

>. 

Q. 

2 


10 
8 
6 
4 
2 
0 


100 


200 


300 


30 


20 


0  W      '  200  500      ^~ 


1500 

1000 

500 

0 

200 
=•  150 

•S  100 

X 

Z   50 

0 

400 
300 
200 
100 


■^J^^^^^ 


100 


200 


^00" 


1 00  200  300 


vtrfiyWMllW,^^^, 


100 


200 


300 


1 

o 


200 
150 
100  [ 


50 


^y*%%»^^ ■ 


Figure 


100  200  300 

6.35  Temporal  water  quality  variations  at  CH010  station  in  2000. 


226 


o 

Q 


3 

(0 

>< 

Q. 

2 
o 

.c 
o 


100 


200 


3. 


2000 

1500 

1000 

500 

0 

400 
300 
200 
100 

0 


600 


100 


200 


;to;*S;*s*^^ 


100 


200 


=;  150 
~  100 

o 

a 


50  ■        _ 

u  100 


300 


300 


300 


200  300 

Figure  6.36  Temporal  water  quality  variations  at  CES02  station  in  2000. 


227 


400  r 

~.     300 

■5. 

3     200 

Q. 

*-     100 

0 

200 


^  150 

«  100 

o 

Q.  50 


100 


200 


300 


4UU 

a 

i  | 

^ 

300 

- 

/P    l/V  ,,| 

?   *V 

eft 

5 

200 

A 

^  l^n/1 

■            " 

I 

*v^             ■ 

A  IM 

Jj  jtitMiiiM 

liiiiAlii  ujAl 

z 

100 

i'-vHi1 'MiWmma 

mnm    11 

iiliiiJlWW      *v^_ 

^^f^y- 

0 

pff^fmPHH 

mf  if 

iilllMi" 

■    V^* 

100 

200 

300 

■ 

■ 

■ 

""^^^             J^^^Jj^^i'''                   | 

■ 

\~s 

100 

200 

300 

jmBS^jmL 


A 

/^        W-v  I 


100 ""  200  300 

Figure  6.37  Temporal  water  quality  variations  at  CES03  station  in  2000. 


228 


| 

E 

O 
Q 


§  10° 

3   80  F 

a 

=   60 


100 


200 


300 


•q.  40 

o 

o  20 

5  o 


"       100  200  %6o 


"ft 

z 


2000 

1500 

1000 

500 

0 

400 
=;  300 


^*«<€Mft 


MPp 


0H 


l^P*»*^»w>B^^ 


100 


200 


300" 


200 


■a. 

2 


0 

400 
300 
200 
100 


100  -^^^^^ 

0  TOO  '       2$0  300 


"-*****, 


WWjW  i  f&HRKlmmim 


100 


illliliWilliiMHjIWIlWM 


200 


gggg^ggggggg  i  ^£»f*>m 


300 


i 

3 

o 

Q. 


Figure 


200 
150 
100  i 

50 
0 


J-« "^ 


100  200  300 

6.38  Temporal  water  quality  variations  at  CES08  station  in  2000. 


229 


10 


s=>    8 
% 

E 

O 

Q 


E    6  "  **to,*J">>**^ 


100 


200 


300 


%       30 


3 

>. 

sz 
a. 
o 


20 


10 


0 


.^""^^frlt)^ 


1500 


^>  1000 


200 


250 


300 


lI       v^^^"^'^^^ifc»lyl^V^***w*^^^l J  .nni'Mi  i**mj 


100 


200 


300 


200  r 


150 
100 

50 


hy^r^V-^^mt^K^^s^^  *V*^ 


fti**mH**\nrt>,"«'*r'iH/(f 


100 


200 


300 


3. 


200  r 
150 


100 

50 


i 

o 

a 


0 

200 
150 
100 


H^4w»"*^^ 


100 


200 


300 


50 


0  '   rr 


Figure 


100  200  300 

6.39  Temporal  water  quality  variations  at  CHOI 3  station  in  2000. 


230 


"e 

(D 


O 

.2 
Q 

> 


35 


g       30 

Q. 

£       25 


W        20 


15 


Surface 

Bottom 

*¥*V 

50 


100 


150 


200 


250 


300 


O 

E 

re 
k_ 
0) 

a 

E 

0) 


•S      15 


re 

E 

S 

c 
o 

E 
a> 
re 
v 
O. 


1> 

s 

c 


T3 
1 

o 
in 


re 

"e 

i 

Q 
O 
C/3 


1 50  200 

Julian  Day 


300 


Figure  6 
possible 
fluxes  at 


.40  The  comparison  between  simulated  dissolved  oxygen  concentration  and  the 
causes  of  hypoxia:  river  discharge,  salinity,  temperature,  and  re-aeration  and  SOD 
CH006  water  quality  measurement  station. 


231 


6/18/2000  13:00  (Julian  Day  170) 
CH006  CH005  CH004 


-22500     -20000     -17500     -15000     -12500 

Distance  from  Peace  River  upstream  (m) 


10000 


>5000 


CH006 


CH005 


CH004 


-22500  -20000  -17500  -15000  -12500 

Distance  from  Peace  River  upstream  (m) 


-10000 


Figure  6.41  Simulated  longitudinal-vertical  salinity  and  dissolved  oxygen  concentration 
along  the  Peace  River  at  1  pm  on  June  18  (Julian  Day  170),  2000. 


232 


10/6/2000  13:00  (Julian  Day  280) 

CH005  CH004 


-22500     -20000     -17500     -15000     -12500 

Distance  from  Peace  River  upstream  (m) 


10000 


25000 


-22500     -20000     -17500     -15000     -12500 

Distance  from  Peace  River  upstream  (m) 


10000 


Figure  6.42  Simulated  longitudinal-vertical  salinity  and  dissolved  oxygen  concentration 
along  the  Peace  River  at  1  pm  on  October  6  (Julian  Day  280),  2000. 


233 


Chlorophyll  a  (p.g/L) 


Chlorophyll  a  (ug/L) 


ft  •  ^ 


Chlorophyll  a  (ug/L) 


Chlorophyll  a  (ug/L) 


Figure  6.43    Simulated  near-surface  chlorophyll_a  concentration  in  Charlotte  Harbor 
estuarine  system  on  February  9,  May  9,  August  7,  and  November  5,  2000. 


234 


Near-Bottom  Dissolved  Oxygen  (mg/L) 


Near-Bottom  Dissolved  Oxygen  (mg/L) 


V  \  .  > 


$e 


Near-Bottom  Dissolved  Oxygen  (mg/L) 


Near-Bottom  Dissolved  Oxygen  (mg/L) 


August  7,  2000 


I 


November  5,  2000 


&*$■%! 


J.  tAa? 


Bvl 


J 


^ 


Figure  6.44  Simulated  near4Dottom  dissolved  oxygen  concentration  in  Charlotte  Harbor 
estuarine  system  on  February  9,  May  9,  August  7,  and  November  5,  2000. 


v 


235 
6.4.3  Application  of  2000  Water  Quality  Simulations 

Hydrologic  alteration 

The  hydrodynamic  and  water  quality  models  of  the  Charlotte  Harbor  estuarine  system 
have  been  successfully  developed  and  validated.  The  models  can  be  used  to  address  the 
effect  of  hydrologic  alteration  on  the  water  quality  as  shown  in  Chapter  5.  Using  the 
validated  water  quality  model,  we  performed  several  model  simulations  with  the  causeway 
islands  removed  and  with  the  IntraCoastal  Waterway  removed  during  April  9  to  June  10, 
2000.  The  results  are  compared  with  those  under  existing  conditions  with  both  of  them  in 
place.  The  scenarios  and  stations  for  comparison  are  the  same  as  those  for  comparison  of 
flow  and  salinity  in  Chapter  5. 

The  chlorophyll_a  concentrations  for  both  cases  were  compared  with  those  for  the 
baseline  simulation  in  Figure  6.45.  The  results  at  all  three  stations  show  that  chlorophyll_a 
concentration  is  not  noticeably  affected  by  the  absence  of  Intracoastal  Waterway  (NICW 
case)  or  the  causeway  (NSC  case).  Just  like  the  chlorophyll_a  concentration,  the  other  water 
quality  species  do  not  show  much  effect  by  these  hydrologic  alterations. 

To  quantify  the  effect  of  these  hydrologic  alterations  on  the  spatial  distribution  of 
water  quality  species,  the  snapshots  of  chlorophyll_a  and  dissolved  ammonium  nitrogen 
concentrations  for  these  two  cases  were  compared  with  those  for  the  baseline  simulation  in 
Figures  6.46  and  6.47.  There  is  not  noticeable  impact  on  the  water  quality  species 
distribution  in  the  San  Carlos  Bay  area,  consistent  with  the  negligible  effect  of  hydrologic 
alteration  on  flow  and  salinity  in  Chapter  5. 

Table  6.6  shows  the  temporal  average  water  quality  species  concentrations  at  the 
eleven  comparison  stations  for  the  baseline  run.  Results  of  the  hydrologic  alteration  cases 


236 
were  compared  to  the  baseline  results  and  RMS  differences  calculated,  then  normalized  with 
these  averaged  water  quality  species  concentration  in  Table  6.6.  Normalized  RMS 
differences  of  water  quality  species  concentrations  for  the  no  causeway  case  from  April  to 
June,  2000  are  shown  in  Table  6.7.  Table  6.8  shows  the  normalized  RMS  differences  for  the 
no  ICW  case,  while  Table  6.9  shows  the  normalized  RMS  differences  when  both  the 
causeway  islands  and  the  ICW  are  removed.  The  RMS  differences  are  less  than  2%  at  all 
selected  stations  for  hydrologic  alteration.  Therefore,  it  can  be  concluded  that  neither  the 
causeway  islands  nor  the  Intracoastal  Waterway  had  noticeable  effect  on  the  water  quality 
in  the  San  Carlos  Bay  and  Pine  Island  Sound  area. 


237 


ST05 

BASELINE 

30 

Causeway  removed  (NSC) 

i25 

Intra  Coastal  Waterway  removed  (NICW) 

a. 

"-"    20 

CO 

^    15 

£ 

a 

,aAWW/Vu,u 

g    10 

^         ^^w 

o 

,.    AAAA/W 

1                                    ^^,,jy/W\MM/\IVVw^ 

■=        5 

-,^a^A*W^ 

v  V\jW™                                                                      'V'*  " 

O 

1 

II                                  1 

0 

12C 

)                         140.        _              160                         180 

"9> 

CO 


2 
o 


30 


ST08 


25  - 


"a. 

a. 

•— '    20 

CO 

%    15 
Q. 

2    10 
o 

■=        5 
O 

0 


ST10 


120 


120 


140.       _  160 

Julian  Day 


180 


160 


180 


14Q.       _ 
Julian  Day 

Figure  6.45  Comparisons  of  simulated  surface  chlorophyll_a  concentration  for  three  cases 

at  three  selected  stations:  ST05  (Pine  Island  Sound),  ST08  (San  Carlos  Bay),  and  ST  10 

(Caloosahatchee  River  mouth) 


238 


Chlorophyll  a 

20 

15 
10 

5 

•  S  xk 

Figure  6.46  Comparisons  of  simulated  surface  chlorophyll_a  concentration  fields  in  San 
Carlos  Bay  after  90  days  of  simulation  for  three  cases. 


239 


'  m 


Figure  6.47    Comparisons  of  simulated  surface  dissolved  ammonium  nitrogen  (NH4) 
concentration  fields  in  San  Carlos  Bay  after  90  days  simulation  for  three  cases. 


240 


Table  6.6  The  temporally  averaged  water  quality  species  concentrations  for  the  baseline 
2000  simulation. 


Station  number 

ChlA 

DO 

NH4 

NOX 

TKN 

P04 

PHOST 

TOC 

(mg/L) 

(mg/L) 

(u-g/L) 

(Hg/L) 

(Hg/L) 

(Hg/L) 

(lig/L)   ( 

mg/L) 

1 

6.04 

6.39 

99.53 

31.69 

728.98 

9.11 

103.50 

1.61 

2 

6.51 

6.39 

103.06 

29.60 

765.36 

9.62 

105.38 

1.68 

3 

7.28 

6.46 

101.99 

27.46 

763.13 

9.63 

110.97 

1.69 

4 

9.14 

6.50 

103.79 

24.03 

784.51 

9.69 

141.30 

2.07 

5 

7.37 

6.18 

115.69 

26.32 

864.87 

10.60 

90.78 

2.35 

6 

7.59 

6.33 

110.88 

25.46 

839.52 

10.39 

109.21 

1.97 

7 

7.24 

6.40 

107.63 

26.88 

809.50 

10.10 

108.62 

1.79 

8 

8.57 

6.48 

103.77 

22.69 

785.02 

10.35 

116.44 

1.92 

9 

11.10 

6.50 

103.05 

20.03 

787.49 

9.61 

145.89 

2.72 

10 

13.71 

6.44 

101.44 

15.66 

784.27 

9.35 

148.24 

3.85 

11 

10.14 

6.38 

100.42 

16.16 

771.49 

11.44 

117.82 

2.38 

AVG 

8.61 

6.40 

104.66 

24.18 

789.47 

9.99 

118.01 

2.18 

Table  6.7  Normalized  RMS  differences  of  water  quality  species 

concentrations  at  1 1 

stations  during  April  to  June  2000,  i 

showing  the  effect  of  no  causeway  islands. 

Station  number 

ChlA 

DO 

NH4 

NOX 

TKN 

P04 

PHOST 

TOC 

(%) 

(%) 

(%) 

(%) 

(%) 

(%) 

(%) 

(%) 

1 

0.27 

0.25 

0.83 

1.33 

1.36 

0.26 

1.58 

0.24 

2 

0.37 

0.28 

1.17 

1.16 

1.91 

0.29 

1.29 

0.27 

3 

0.65 

0.24 

0.83 

1.56 

1.36 

0.39 

1.9 

0.34 

4 

0.85 

0.24 

1.04 

0.58 

1.77 

0.21 

2.25 

0.31 

5 

0.58 

0.68 

1.35 

1.14 

2.55 

0.45 

2.2 

0.55 

6 

0.47 

0.43 

1.15 

1.19 

2.02 

0.33 

1.95 

0.3 

7 

0.44 

0.32 

1.15 

1.17 

1.95 

0.25 

1.62 

0.23 

8 

0.5 

0.27 

0.94 

1.59 

1.46 

0.4 

2.22 

0.41 

9 

0.93 

0.2 

0.9 

0.41 

1.54 

0.2 

1.93 

0.39 

10 

0.76 

0.18 

0.74 

0.29 

1.23 

0.15 

1.44 

0.36 

11 

0.33 

0.24 

1.09 

1.22 

1.44 

0.28 

1.39 

0.25 

AVG 

0.56 

0.3 

1.02 

1.06 

1.69 

0.29 

1.8 

0.33 

241 

Table  6.8  Normalized  RMS  differences  of  water  quality  species  concentrations  at  1 1 
stations  during  April  to  June  2000,  showing  the  effect  of  no  ICW. 

Station  number    ChlA        DO         NH4        NOX       TKN        P04      PHOST    TOC 

(%)  (%)  (%)  (%)  (%)  (%)  (%)        (%) 


1 

0.28 

0.05 

0.12 

0.42 

0.24 

0.11 

0.46 

0.13 

2 

0.33 

0.08 

0.23 

0.65 

0.38 

0.11 

0.33 

0.19 

3 

0.34 

0.08 

0.18 

0.41 

0.33 

0.22 

0.63 

0.19 

4 

0.97 

0.12 

0.37 

0.75 

0.41 

0.24 

1.89 

0.37 

5 

0.37 

0.30 

0.44 

1.13 

0.75 

0.16 

0.53 

0.49 

6 

0.52 

0.19 

0.39 

1.16 

0.59 

0.16 

0.53 

0.35 

7 

0.46 

0.12 

0.34 

0.95 

0.51 

0.14 

0.56 

0.26 

S 

0.50 

0.17 

0.24 

0.38 

0.47 

0.37 

0.84 

0.30 

9 

1.39 

0.20 

0.62 

0.86 

0.51 

0.29 

2.67 

0.68 

10 

1.51 

0.41 

0.74 

0.82 

0.68 

0.51 

5.31 

2.11 

11 

0.48 

0.13 

0.32 

0.41 

0.56 

0.26 

1.03 

0.41 

AVG         0.65         0.17  0.36         0.72         0.49         0.23         1.34      0.50 

Table  6.9  Normalized  RMS  differences  of  water  quality  species  concentrations  at  1 1 
stations  during  April  to  June  2000,  showing  the  effect  of  no  causeway  island  and  ICW 

Station  number    ChlA        DO         NH4        NOX       TKN        P04      PHOST    TOC 

(%)  (%)  (%)  (%)  (%)  (%)  (%)        (%) 


1 

0.40 

0.27 

0.95 

1.69 

1.36 

0.26 

1.30 

0.30 

2 

0.39 

0.34 

1.39 

1.77 

2.11 

0.30 

1.33 

0.39 

3 

0.72 

0.28 

0.96 

1.89 

1.25 

0.43 

1.46 

0.40 

4 

0.57 

0.24 

1.34 

1.24 

1.47 

0.30 

1.69 

0.39 

5 

0.40 

0.93 

1.77 

2.18 

3.23 

0.54 

2.58 

0.97 

6 

0.42 

0.59 

1.51 

2.22 

2.43 

0.39 

2.19 

0.59 

7 

0.40 

0.42 

1.44 

2.01 

2.23 

0.26 

1.75 

0.42 

8 

0.69 

0.38 

1.03 

1.96 

1.12 

0.48 

2.06 

0.52 

9 

1.07 

0.23 

1.42 

1.15 

1.17 

0.35 

1.80 

0.74 

10 

1.40 

0.35 

1.36 

0.89 

0.75 

0.50 

4.28 

2.19 

11 

0.47 

0.30 

1.26 

1.58 

0.86 

0.37 

0.75 

0.32 

AVG  0.63         0.39  1.31  1.69         1.63         0.38         1.93      0.66 


242 
River  load  reduction 

To  provide  a  tool  for  studying  management  options  and  corresponding  responses  of 
the  Charlotte  Harbor  estuarine  system,  which  is  one  of  the  primary  objectives  of  this  study, 
the  validated  integrated  model  was  used  to  evaluate  the  effectiveness  of  load  reductions  for 
improving  estuarine  water  quality.  To  achieve  this  goal,  it  is  necessary  to  use  a  common  set 
of  initial  conditions  for  the  water  column  and  the  sediments  so  that  any  differences  observed 
between  the  results  of  different  load  reduction  simulations  would  be  attributable  to  the 
differences  in  load  reduction.  To  analyze  the  potential  impact  of  reduced  nutrient  loadings 
to  the  system,  model  simulations  were  carried  out  using  100%  load  reductions  of  nitrogen 
species  concentrations,  100%  load  reductions  of  phosphorous  species  concentrations,  and 
100%  load  reduction  of  nitrogen  and  phosphorous  species  concentrations  at  Peace  River  and 
Caloosahatchee  River. 

One  major  assumption  in  the  load  reduction  runs  is  that  the  SOD  values  are  assumed 
to  be  the  same  as  the  baseline  run.  This  is  because  the  fact  that  SOD  values  for  the  baseline 
run  were  provided  from  field  experiments  and  the  model's  inability  to  directly  related  the 
SOD  values  inside  estuary  to  nutrient  loading  from  the  rivers.  Hence  the  results  presented 
in  the  following  should  be  interpreted  with  caution,  since  large  errors  may  be  associated  with 
the  above  assumption.  In  particular,  since  the  SOD  values  are  unchanged,  load  reduction  is 
not  expected  to  produce  any  improvement  in  hypoxia,  hence  the  DO  results  are  practically 
unchanged  in  the  model  results.  In  reality,  load  reduction  is  expected  to  lead  to  increased 
DO. 

For  the  Peace  River  load  reduction  simulations,  CH004  and  CH006  stations  are 
selected  to  compare  the  water  quality  species  before  and  after  load  reduction.  Figures  6.48 


243 
and  6.49  show  the  water  quality  species  at  CH004  and  CH006  before  and  after  100% 

nitrogen  load  reduction.  Due  to  reduction  of  nitrogen  loading,  the  chlorophyll_a 
concentration  decreased  at  both  stations  because  the  reduction  of  dissolved  ammonium 
nitrogen  as  the  limiting  factor  for  phytoplankton.  The  amount  of  the  decrease  was  12  |ig/L 
at  the  CH004  station  and  4  [igfL  at  the  CH006  station.  The  dissolved  oxygen  concentration 
is  little  decreased  at  the  CH004  station  because  the  amount  of  dissolved  oxygen  from 
photosynthesis  is  reduced  due  to  decreasing  of  phytoplankton  biomass,  while  those  from  re- 
aeration  and  SOD  fluxes  remained  the  same  as  the  baseline  simulation. 

Figures  6.50  and  6.51  show  the  water  quality  species  at  CH004  and  CH006  before 
and  after  100%  phosphorous  load  reduction  from  the  Peace  River.  The  chlorophyll_a 
concentrations  decreased  at  both  stations.  The  phytoplankton  food  limiting  factor  is  changed 
to  soluble  reactive  phosphorous  after  SRP  concentration  reached  to  less  than  phosphorous 
half  saturation  rate.  The  amounts  of  decrease  8  [ig/L  at  CH004  station  and  4  |ig/L  at  CH006 
station  were  smaller  than  those  for  100%  nitrogen  load  reductions.  Dissolved  ammonium 
nitrogen  concentration  was  increased  because  phytoplankton  consumed  a  smaller  amount  of 
NH4  concentration,  as  phytoplankton  food  limit  was  decreased  by  SRP  concentration. 

For  the  Caloosahatchee  River  load  reduction,  CES02  and  CES08  stations  are  selected 
to  compare  the  water  quality  species  before  and  after  load  reduction.  Figures  6.52  to  6.53 
show  the  water  quality  species  at  CES02  and  CES08  before  and  after  100%  nitrogen  load 
reduction,  while  Figures  6.54  to  6.55  show  the  corresponding  results  for  100%  phosphorous 
load  reduction.  With  reduction  of  nitrogen  or  phosphorous  loading,  the  chlorophyll_a 
concentration  decreased  at  CES02  while  there  is  no  significant  difference  at  CES08.  The 
Chlorophyll_a  concentration  at  CES02  reached  94  |ig/L  due  to  large  amount  of  chlorophyll_a 


244 
from  the  river  loading.  Most  of  the  water  quality  species  are  strongly  affected  by  the  loading 
from  Caloosahatchee  River,  as  shown  in  the  results  at  these  two  stations.  The  impacts 
resulting  from  load  reductions  were  confined  in  Caloosahatchee  River  and  became 
insignificant  once  outside  the  Caloosahatchee  River. 

As  mentioned  earlier,  dissolved  oxygen  concentration  was  relatively  unaffected  by 
the  loading  reduction  because  the  SOD  kinetic  process  used  in  water  quality  model  is  an 
empirical  formula  which  is  a  function  of  temperature,  dissolved  oxygen,  and  sediment  type. 
Sediment  oxygen  demand  (SOD)  depends  on  the  deposition  and  decomposition  of  organic 
matter  on  the  seabed,  and  the  exchange  of  nutrients  and  oxygen  across  the  sediment-water 
interface.  To  provide  detailed  trends  in  hypoxia  in  response  to  organic  loads,  it  is  necessary 
to  apply  sediment  flux  model  (DiToro  and  Fitzpatrick,  1993)  with  observed  data  for  CH4, 
H2S,  and  organic  matter  in  sediment  column  and  river  boundary  as  described  in  Figure  4.4. 
Although  applying  of  sediment  flux  model  provides  rational  predictions  of  sediment  response 
to  environmental  alterations,  it  requires  additional  information  as  compared  to  the  use  of 
empirical  SOD  model. 

Without  any  available  data,  the  sediment  flux  model  was  tested  for  organic  matter 
river  load  reduction  in  Peace  River  using  DiToro's  method  described  in  Appendix  F.  Figure 
6.56  shows  the  comparison  of  chlorophyll_a  and  dissolved  oxygen  concentrations  at  CH006 
station  between  baseline  and  100%  organic  matter  loading  reduction  simulations.  The  result 
does  not  show  much  difference  in  dissolved  oxygen  concentrations  between  the  baseline  and 
100%  organic  matter  load  reduction  simulations,  because  SOD  flux  from  methane  and 
nitrogen  calculated  by  the  sediment  flux  model  is  not  enough  to  create  hypoxia  in  CH006. 
The  measured  total  organic  carbon  concentration  at  CH006  is  much  lower  than  that  at  the 


245 
Peace  River  boundary  (CH029).  Therefore,  the  organic  carbon  concentration  from  Peace 
River  does  not  reach  to  the  hypoxia  area  (CH006).  In  the  summer  season,  the  dissolved 
oxygen  concentration  of  the  river  load  reduction  scenario  is  even  lower  than  that  for  the 
baseline  simulation  due  to  reduced  chlorophyll_a  concentration.  Chlorophyll_a 
concentration  from  river  loading  in  summer  season  is  quickly  reduced  because  of  the  nutrient 
load  reduction  from  the  river.  The  photosynthesis  is  decreased  due  to  chlorophyll_a 
decrease,  but  CBOD  is  increased  in  the  water  column  due  to  mortality  of  phytoplankton  and 
settling  to  the  sediment  layer.  This  increase  in  CBOD  will  increase  the  SOD  flux  at  the 
sediment-water  interface. 

The  measured  data  show  that  total  organic  carbon  concentration  at  CH006  was  very 
low  during  the  simulation  period.  Hence,  the  carbonaceous  oxygen  demand  may  not  be  a 
major  factor  of  SOD  flux  in  upper  Charlotte  Harbor.  Sulfide  flux  could  be  a  very  important 
component  of  SOD  in  anoxic  estuarine  water  (Chapra,  1997)  such  as  the  upper  Charlotte 
Harbor.  Hence,  sulfide  and  iron  fluxes  should  be  included  in  the  sediment  flux  model  to 
better  represent  sediment  oxygen  demand  for  river  load  reduction  simulations.  In  addition, 
more  field  data  and  modeling  effort  to  determine  sediment  fluxes  should  be  focused  on  the 
specific  conditions  of  upper  Charlotte  Harbor,  especially  specifying  the  parameters  for  the 
sediment  layer  and  the  sediment- water  interface. 

To  test  the  effect  of  reducing  organic  matter  in  river  loading  with  current  water 
quality  model,  the  analysis  of  the  potential  impact  of  reduced  organic  matter  loadings  to  the 
system  was  carried  out  by  model  simulations  using  50%  and  75%  SOD  load  reductions  at 
Peace  River.  Figure  6.57  shows  the  comparison  of  dissolved  oxygen  concentrations  at 
CH004  and  CH006  before  and  after  50%  SOD  reduction  and  75%  SOD  reduction.  The 


246 
dissolved  oxygen  concentration  in  the  upper  Charlotte  Harbor  is  very  sensitive  with  SOD 

reduction.    With  50%  SOD  load  reduction,  the  surface  and  bottom  dissolved  oxygen 

concentration  are  increased  from  maximum  2  mg/L  to  0.2  mg/L.  With  75%  load  reduction, 

no  hypoxic  event  was  found  during  this  simulation  and  the  dissolved  oxygen  distribution 

exhibit  minimal  vertical  stratification.  Although  there  is  no  hypoxia  event,  some  localized 

low-levels  of  near-bottom  dissolved  oxygen  and  vertical  stratification  were  maintained  even 

during  the  75%  SOD  load  reduction  scenario. 

To  quantify  the  relationship  between  SOD  coefficient  and  hypoxia,  the  area  which 
exposed  in  bottom  water  hypoxia  condition  (DO  is  less  than  2  mg/L)  was  calculated  with  0.5 
(75%  reduction),  1.0  (50%  reduction),  and  2.0  (baseline)  rate  constant  of  SOD  at  20  °C 
during  the  simulation  period  (Figure  6.56).  With  2.0  as  the  rate  constant  of  SOD,  over  60 
km2  of  the  upper  Charlotte  Harbor  was  hypoxic  during  the  summer  season,  while  hypoxia 
condition  was  not  observed  with  0.5  as  the  SOD  rate  constant.  The  maximum  areal  extent 
of  hypoxia  67  km2  was  on  August  21,  2000.  This  was  also  accompanied  by  a  high  degree 
of  stratification  as  indicated  in  Figure  6.40. 

With  further  scientific  understanding,  the  integrated  modeling  approach  would  enable 
the  development  of  a  science-based  management  tool  which  is  built  on  process-based 
understanding  rather  than  simple  regression.  Subsequent  refinement  of  this  integrated  model 
can  be  used  to  address  ecosystem  management  issues  such  as  controlling  estuarine 
eutrophication  and  determining  allowable  external  nutrient  loading  levels  to  restore  water 
quality  in  estuarine  system. 


247 


10 

.— - 

8 

E 

6 

•^^ 

O 

4 

Q 

2 

°, 

1 

50 

3 

40 

(D 

^T 

30 

>. 

SZ 
Q. 

20 

O 

O 

10 

SZ 

O 

U, 

260 


280 


baseline  near  surface 

baseline  near  bottom 
-100  %  N  load  reduction  near  surface 
'  1 00  %  N  load  redcction  near  bottom 


VA*vvw*w*s#//w\tf*MMwWA 


60 


2000 


200 


^  150 

*  100 


50 


i60 


180 


200 


220 


240 


260 


280 


260 


280 


180 


200 


ijiiiiliiijiiiljjiiiiiiiiliiiiiiiiiii 
220 


i 


dlWilillLi    '■'■ 


liiuilVai'i'iili'uiiiiii'ii'iVuimiii 


240 


260 


280 


300 


320 


300 


320 


300 


320 


300 


320 


ui60      180      200      220      240      260      280      300    ~ 320 

Figure  6.48   The  water  quality  species  concentrations  at  CH004  water  quality  measured 
station  before  and  after  100  %  nitrogen  load  reduction  from  the  Peace  River. 


248 


10 

^^ 

8 

E 

6 

O 

4 

D 

2 

°, 

1 

50 

3 

40 

CO 

— 

30 

>> 

a. 

20 

o 

-  ***$* '  SJrS 


60 


180 


200 


'^%M^r% 


220 


240 


'fi^ 


260 


280 


10 


o 


baseline  near  surface 

baseline  near  bottom 

100  %  N  load  reduction  near  surface 

1 00  %  N  load  redcction  near  bottom 


^VWtilii^^HWWi 


1 60 


180 


200 


220 


240 


260 


280 


200 


220 


240 


260 


280 


400 

i 

s 

300 
200 

Q. 

100 

Q 


300 


320 


300 


320 


320 


300 


320 


160 


180 


Tor 


220 


240 


260 


280 


300 


320 


1 

O 

Q. 


200 
150 
100 

50 
Q 


V^. 


Figure 
station 


160      180      200      220      240      260      280      300      320 

6.49   The  water  quality  species  concentrations  at  CH006  water  quality  measured 
before  and  after  100  %  nitrogen  load  reduction  from  the  Peace  River. 


249 


5 

o 

Q 


■"TW  ^ 


ui60 


ML  .  \^x^  %^jM'ku^..jf  II 

200      220      240      260      280 


1  '•».., 


*% 


50 
40 

m 

=       30 
q.       20 

2 

o 


10 


o 


180 


baseline  near  surface 

baseline  near  bottom 

100  %  N  load  reduction  near  surface 

1 00  %  N  load  redcction  near  bottom 


300 


320 


MwtomHttmi&NNtMNHHNlkNItlllt 


160 


180 


200 


"221T 


240 


260 


280 


300 


320 


320 


300 


320 


1000 

800 

"j*  600 

a   400  I 

200 


ui60 

500 
e.  400  r 
2  300 


mftmt^mt^^ 


180 


200 


220 


240 


260 


280 


300 


320 


O 
a 


200 
100 

q 


^*^m*kmiiiML 


Hm- 


m^^^Mj:^lm^ 


iMMftto 


1 60      180      200      220      240      260      280      300      320 

Figure  6.50  The  water  quality  species  concentrations  at  CH004  water  quality  measured 
station  before  and  after  100  %  phosphorous  load  reduction  from  the  Peace  River. 


250 


o 


10 
8 
6 
4 
2 
Q 


'^^^^m^mL^' 


ML3 


W^^ 


60 


180 


200 


220 


240 


260 


280 


300 


320 


*>  50 

3  40 
n 

=  30 

q.  20 
o 

S  10 

5  °.6ir 

2000 


fi  1500 
w  1000  l 

z 

*      500 


baseline  near  surface 

baseline  near  bottom 
- 1 00  %  N  load  reduction  near  surface 
'  1 00  %  N  load  redcction  near  bottom 


^*Vftnt»»»»» 


160 


m  mmmm^0^Smmmmm0S»j 


ttUUtU  MttHUUk,,,   ,/(|i 


180      200 


220 


240      260 


280 


300 


320 


•""'••vmutHMinmii 


4 -»»»» 


1000 

• 

800 

. 

1>  600 

. 

i 

q.  400 

■ 

H 

200 

180 


200 


<G&m0Baaai£m^kmu* 


220 


240 


260 


320 


o 

Q. 

SUU 

400 
300 
200 
100 

■ 

..^..^Wf^few.-. 

4* 

>0 

180 

200 

220      240 

260"     280      360      32 

Figure  6.51    The  water  quality  species  concentrations  at  CH006  water  quality  measured 
station  before  and  after  100  %  phosphorous  load  reduction  from  the  Peace  River. 


251 


100 


baseline  near  surface 

baseline  near  bottom 

1 00  %  N  load  reduction  near  surface 

1 00  %  N  load  redcction  near  bottom 


400 
^     300 

*     200 

X 

Z      100 


•     J 

i 

tyi|i||jjijj|tj|j ' 

ij|f|fUj|L,            | 

*'i 

60 

180 

200 

220              240 

260 


280 


300 


320 


ui60  180  200  220  240  260  280  300  320 

Figure  6.52   The  water  quality  species  concentrations  at  CES02  water  quality  measured 
station  before  and  after  100  %  nitrogen  load  reduction  from  Caloosahatchee  River. 


252 


■a 

E 

O 

D 


>> 
SZ 
D. 
O 


10 
8 
6 
4 
2 


MWWWIflMWM 


w^w *^«HWtlHw»w^^ 


Hmm+tw+^H 


'i60 


180 


200 


220 


240 


260 


280 


300 


320 


50 
40 
30 
20 

10  f- 


baseline  near  surface 
baseline  near  bottom 
- 1 00  %  N  load  reduction  near  surface 
1 00  %  N  load  redcction  near  bottom 


1**M*w<*nv™,w/A^/llfl 


WWWMlwilWHHi^ 


60 


180 


200 


220 


240 


260 


280 


300 


320 


3 
z 
i- 


2000 

1500 

1000 

500 


^m^mm#mtumm\w* 


i60 


180 


200 


220 


240 


260 


280 


300 


320 


5 

a. 


400 
300 
200 
100 


^^m^^^^^S^^^SSZ 


w*"M**Hvw'wu 


ui60 
400  r 
300 
200 
100 


180 


200 


220 


240 


260 


280 


300 


320 


ikMHUuHHnttnmttm"  >tf* m  '*  "■«*" ***mfiWtm*nM  '"*%l»M»%^|/r^^ 


1 

O 

Q. 


ui60 
200  r 
150 
100 

50 


180 


200 


220 


240 


260 


"280* 


300 


*320 


0 


MHHttttt0lto*i  *  i  WMUMHH  w 


.ji^fcittii 


i 60      180      200      220      240      260      280      300      320 

Figure  6.53   The  water  quality  species  concentrations  at  CES08  water  quality  measured 
station  before  and  after  100  %  nitrogen  load  reduction  from  Caloosahatchee  River. 


253 


10 

0  8 

E 

O 

Q 


_: 

.ll/V  I 

llnJn  ' 

60 

180 

200 

220             240             260              280              300             32 

baseline  near  surface 

baseline  near  bottom 

1 00  %  N  load  reduction  near  surface 

1 00  %  N  load  redcction  near  bottom 


1000 
800  r 
"^     600 
a      400 

200  r 


0 


,«"•-'" 


"*    ^  **W. 


1 60 


180 


200 


220 


240 


260 


280 


300 


320 


500 
£.  400 
J     300 

J    200 

°-     100 


A, 


,  jfcyTf , ,  fa , 


.v'.':' 


sj^%i 


#T'W| 


W  W  ""%** 


*~J .»— «<ft*»W»»H»Vm~- 


ui60  180  200  220  240  260  280  300  320 

Figure  6.54   The  water  quality  species  concentrations  at  CES02  water  quality  measured 
station  before  and  after  100  %  phosphorous  load  reduction  from  Caloosahatchee  River. 


254 


E 
o 

Q 


=1 


a 

2 
o 

O 


10 

8 

4 
2  f 


immiiimwmw ww 


240 


260 


280 


300 


320 


50 
40  r 

30 
20 
10 


60 


•5. 

a. 


2000 

1500 

1000 

500 


%*. — 


1 60 


"a. 

3 


400 
300 
200 
100 
Cj 

400 
300 
200 
100 


60 


1 60 


o 

Ql 


200 
150 
100 

50 


baseline  near  surface 

baseline  near  bottom 

100  %  N  load  reduction  near  surface 

1 00  %  N  load  redcction  near  bottom 


*— iww»^».M<>(#ffWW^NW^MMiw*<»*w«tMwwMMM1 — riff**^  jt^'-^^^gjij'^i^ViIfy^***'  Will  I'f^lWf'^TyVl'iW'VttUi.u 


180 


200 


220 


240 


260 


280 


300 


320 


^^iimmmm^''fmtmm^ 


1  ■   'numwmwm 


180 


200 


"22T 


240 


260 


280 


300 


320 


180 


200 


"22T 


240 


260 


~2&0 


300 


320 


HH^Min^llll^^ 


180 


200 


220 


240 


260 


280 


300 


320 


IWMXIMilitJl  fcWWlWWMMMM 


,mi, ).-v,'i,"i-,-,--^--l'-;ri,,',;"'  ■"■  ■■"■->-- 


ui60      180      200      220      240      260      280      300      320 

Figure  6.55   The  water  quality  species  concentrations  at  CES08  water  quality  measured 
station  before  and  after  100  %  phosphorous  load  reduction  from  Caloosahatchee  River. 


255 


50 


100 


150  200  250 

Julian  Day 


300 


10r 


c 

D) 


I 

O 


8 


0 


50 


at  surface  for  Baseline  simulation 

at  bottom  for  Baseline  simulation 

at  surface  for  1 00  %  organic  matter  reduction 

at  bottom  for  1 00  %  organic  matter  reduction 


100 


250 


-I I I !_ 


300 


i    i    i    i 


150  200 

Julian  Day 

Figure  6.56  Dissolved  oxygen  and  chlorophyll_a  concentrations  at  CH006  water  quality 
measured  station  before  and  after  100  %  organic  matter  load  reduction  from  Peace  River 
using  DiToro's  sediment  flux  model. 


256 


50  %  SOD  reduction  at  CH004  station 


1 
E 

c 

<D 
U> 
>• 
X 

O 

■a 

I 

o 
en 
w 


6 

c 
>< 

X 

O 

J 

O 

(/> 

W 


§ 

c 
a) 

5. 
x 

O 

-o 

I 
o 
</> 


ian  Day 
50  %  SOD  reduction  at  CH006  station 


10r 

8 

6 
4 
2 
0 


luflan  Day  200  250  300 

75  %  SOD  reduction  at  CH004  station 


Surface,  for  Baseline 
Bottom,  for  Baseline 
Surface,  for  Reduction 
Bottom,  for  Reduction 


m 


ian  Day 
75  %  SOD  reduction  at  CH006  station 


irian  Day 

Figure  6.57  Dissolved  oxygen  concentrations  at  CH004  and  CH006  water  quality 
measured  stations  before  and  after  50  %  SOD  reduction  and  75%  SOD  reduction. 


257 


80  r 


70 


eg 

E 

60 

-* 

(0 

50 

<D 

l_ 

< 

40 

A3 

X 

o 

30 

Q. 

> 
X 

20 

10 

0 

SOD  =  2.0  g/m3/day 

SOD  =  1.0g/m3/day 
-  -  SOD  =  0.5  g/m3/day 


150  200 

Julian  Day 

Figure  6.58  The  comparison  of  hypoxia  area  at  Upper  Charlotte  Harbor  according  to 
varying  SOD  constant  rate  at  20°C. 


CHAPTER  7 
CONCLUSION  AND  DISCUSSION 

An  integrated  modeling  system,  CH3D-HMS  (Sheng  et  al.,2002),  which  includes  a  3- 
D  hydrodynamics  model,  a  3-D  sediment  transport  model,  and  a  3-D  water  quality  model, 
has  been  enhanced  and  applied  to  the  Charlotte  Harbor  estuarine  system.  Circulation  and 
water  quality  in  the  entire  estuarine  system  have  been  simulated  and  validated  with  field  data 
from  numerous  sources  in  1986,  1996,  and  2000.  To  reproduce  bottom  water  hypoxia  in 
upper  Charlotte  Harbor,  models  of  oxygen  balance  and  oxygen  fluxes  at  the  air-sea  and 
sediment-water  interfaces  are  enhanced.  To  achieve  a  better  understanding  of  the  temporal 
and  spatial  variation  of  temperature  and  light,  and  their  effect  on  water  quality  processes, 
the  3-D  temperature  model,  with  a  heat  flux  model  at  the  air-sea  interface,  and  the  physics- 
based  light  attenuation  model  (Christian  and  Sheng,  2003)  were  applied.  Major  conclusions 
of  this  study  are  summarized  in  the  following: 

1)  A  fine-resolution  numerical  grid  which  accurately  represents  the  complex 
geometrical  and  bathymetric  features  in  Charlotte  Harbor  estuarine  system  was  generated  and 
used  to  simulate  the  hydrodynamic,  salinity  and  temperature  characteristics  in  1986  and 
2000.  A  sensitivity  study  was  conducted  to  examine  the  sensitivity  of  model  results  to  such 
factors  as  boundary  conditions,  model  coefficients,  bathymetry,  advection  scheme,  and  grid 
resolution.  Flow,  salinity  and  temperature  patterns  produced  by  hydrodynamic  model  agree 
well  with  existing  data.  The  normalized  RMS  error  analysis  demonstrated  model's  ability 


258 


259 
to  simulate  water  level,  currents,  salinity  and  temperature  within  7.5  %,  20%,  6.5  %  and  7% 
accuracy,  respectively. 

2)  The  calibrated  circulation  model  was  applied  to  assess  the  impact  of  the  removal 
of  the  Sanibel  Causeway  and  IntraCoastal  Waterway  on  the  flow  and  salinity  pattern  in  the 
San  Carlos  Bay  and  Pine  Island  Sound  and  to  develop  minimum  flow  criteria  for  the 
Caloosahatchee  River.  The  results  show  that  these  hydrologic  alterations  do  not  appear  to 
show  noticeable  impact  on  the  flow  and  salinity  patterns  in  the  San  Carlos  Bay  and  Pine 
Island  Sound.  The  salinity  at  the  Caloosahatchee  River  Mouth  is  reduced  by  about  1.36  ppt 
in  the  absence  of  the  IntraCoastal  Water  way. 

3)  The  relationship  between  salinity  at  Fort  Myers  and  river  discharge  was  established 
by  comparing  1  -day  and  30-day  averaged  salinity  values  at  Fort  Myers  and  fresh  water  inflow 
at  S79  (Figure  5.35).  According  to  this  relationship,  a  total  river  discharge  of  18  m3/s  at  S79 
produces  a  1-day  averaged  salinity  of  20  ppt  at  Fort  Myers. 

4)  The  newly  enhanced  integrated  model,  CH3D-IMS,  was  used  to  perform  1996  and 
2000  hydrodynamic,  sediment  transport,  and  water  quality  simulations  of  the  Charlotte 
Harbor  estuarine  system.  The  normalized  RMS  errors  of  model  simulation  vary  generally 
within  10  -  45%  for  all  stations  and  all  species.  The  agreement  between  simulated  and 
measured  DO  is  particularly  good  with  less  than  20%  error  everywhere.  The  daily  and 
seasonal  vertical  distribution  and  fluctuation  of  DO  are  successfully  simulated  by  applying 
measured  SOD  at  the  upper  Charlotte  Harbor. 

5)  The  upper  Charlotte  Harbor  system  has  been  suffering  summer  hypoxia  in  bottom 
water  in  the  past  decade.  In  this  study,  CH3D-IMS  was  enhanced  and  calibrated  to  analyze 
this  phenomenon.  The  calibrated  model  was  used  to  examine  the  dynamics  of  various  factors 


260 
that  can  affect  hypoxia,  including  freshwater  inflow,  tides,  SOD,  and  water  column  oxygen 

consumption.  The  model  results  suggest  that  hypoxia  in  the  upper  Charlotte  Harbor  is 

primarily  caused  by  the  combined  effects  of  SOD  and  vertical  salinity  stratification,  and  also 

enhanced  by  water  column  oxygen  consumption. 

6)  The  integrated  model  was  applied  to  assess  the  effects  of  hydrologic  alterations 
and  to  provide  a  preliminary  evaluation  of  pollutant  load  reduction  goal  (PLRG).  The  results 
show  that  these  hydrologic  alterations  do  not  appear  to  show  noticeable  impact  on  the  water 
quality  patterns  in  the  San  Carlos  Bay  and  Pine  Island  Sound.  The  dissolved  oxygen 
concentration  of  the  estuarine  system  does  not  response  with  different  load  reduction  because 
of  the  use  of  the  same  SOD  values  for  all  load  reduction  simulation.  In  reality,  SOD  is 
expected  to  be  reduced  when  nutrient  loading  is  reduced.  However,  due  to  lack  of  detailed 
data  of  the  sediment  diagenesis  and  lack  of  understanding  of  the  relationship  between  SOD 
and  nutrient  loading,  the  model  cannot  simulate  the  response  of  SOD  to  load  reduction. 
Further  research  is  needed  to  develop  the  ability  to  simulate  the  effect  of  nutrient  load 
reduction. 

7)  A  systematic  calibration  procedure  has  been  developed  for  a  more  efficient  and 
more  objective  calibration  of  the  water  quality  model.  A  consistent  framework  for 
systematic  calibration  is  formulated  which  include  following  step:  model  initialization, 
sensitivity  analysis,  parameterization,  and  formulation  of  calibration  criteria. 

8)  Although  the  integrated  model  developed  and  applied  in  this  study  have  performed 
well  simulating  hydrodynamic  and  water  quality  components  in  the  Charlotte  Harbor 
estuarine  system,  the  model  contains  various  uncertainties,  assumptions  and  simplifications 
which  need  further  investigation  and  improvement.  Several  possible  improvements  in  the 


261 
developed  integrated  model  and  year-long  simulation  of  hydrodynamic  and  water  quality 
dynamics  in  the  system  are: 

a)  A  finer  grid  system  is  needed  to  represent  the  navigation  channel,  causeway,  small 
islands,  and  complex  shoreline.  With  grid  resolution  used  in  this  study,  the  model  cannot 
solve  the  dynamics  of  the  shallow  regions  less  than  1.5  m  in  depth  of  the  system,  where  wave 
interactions  become  important. 

b)  Sediment  oxygen  demand  (SOD)  depends  on  the  deposition  and  decomposition 
of  organic  matter  on  the  seabed,  and  the  exchange  of  nutrients  and  oxygen  across  the 
sediment-water  interface.  In  order  to  predict  trends  in  hypoxia  in  response  to  organic  loads, 
it  is  necessary  to  include  oxidation-reduction  processes  with  CH4,  H2S,  and  organic  matter 
(DiToroetal.  1990). 

c)  More  accurate  river  discharge  data  are  needed.  The  river  discharge  data  at  Peace 
and  Myakka  rivers  are  underestimated  since  not  all  freshwater  inflows  are  accounted  for  by 
the  data.  To  solve  this  problem,  the  integrated  model  need  to  be  coupled  with  watershed 
model  to  supply  missing  freshwater  discharge 

d)  Phytoplankton  dynamics  are  different  with  different  phytoplankton  species.  To 
better  understand  phytoplankton  process  and  eutrophication  characteristics,  gathering  more 
phytoplankton  species  data  and  applying  models  of  multiple  phytoplankton  species  are 
needed 

e)  Submarine  groundwater  is  an  important  mechanism  in  delivering  chemical 
elements  to  surface  waters,  and  can  be  of  the  same  order  of  magnitude  as  surface  water 
sources  (Oliveria  et  al,  2003).  Rutkowski  et  al.  (1999)  suggest  that  the  nutrient  flux  by 
submarine  groundwater  is  on  the  same  order  of  magnitude  as  a  small  river.  To  address  this 


262 
issue,  a  nutrient  advection  by  coupling  with  groundwater  model  with  the  integrated  model. 

f)  Spatially  more  detailed  and  temporally  more  frequent  hydrodynamic  and  water 
quality  data  inside  the  Charlotte  Harbor  and  at  the  river  mouths  are  needed.  Data  of  nutrients 
and  organic  material  in  the  sediment  column  should  be  gathered  to  enable  simulation  of 
hypoxia  in  the  upper  Charlotte  Harbor. 


APPENDIX  A 
FLOWCHARTS  FOR  CH3D-IMS 

The  following  flow  charts  illustrate  the  order  of  operations  performed  in  the 

integrated  modeling  for  Charlotte  Harbor  estuarine  system.  Figures  A.  1  illustrate  the  flow 

chart  of  the  main  CH3D  program  (ch3d.f).    Main  program,  ch3d.f,  call  the  subroutine 

ch3dm2.f  which  is  the  driving  subroutine  for  time  stepping  of  the  solution  as  well  as  defining 

the  time  varying  forcing  functions  and  generating  output  shown  in  Figure  A.2.  Water  quality 

model  and  sediment  transport  model  are  coupled  in  this  subroutine.   Figure  A.3  and  A.4 

show  the  initializing  sediment  parameters  and  main  routine  of  sediment  transport  model, 

respectively.  Water  quality  model  was  illustrated  in  Figure  A.5  through  A.7. 


263 


Chart  1:  main  program  (ch3d.f) 


264 


Set  screen  output  destination 
(ch3dfileopen.f) 


Output  the  version  number  of  the  fortran/script  files 
(ch3d_versionsf) 


Read  optional  command  line  arguments 
(ch3dclo.f) 


Read  grid  &  bathymetry 

(ch3dgdep  f) 


Read  hydrodynamic  input  file 
(ch3dir_hyd.f) 


Read  salinity  input  file 
(ch3dir_salf) 


Curvilinear  coordinate  transformation  constants 
(ch3dtr.f) 


Modify  bathymetry  (hmin  and  hadd) 
(ch3dih.f) 


Nondimensionalize  and  set  constants 
(ch3dnd.f) 


Set  cell  boundary  direction  flag  and  sweeping  indices 
(ch3dii.f) 


Read  temperature  input  file 
(ch3dir_tem.f) 


Figure  A.l  Flow  chart  for  the  main  CH3D  program. 


265 


Initialize  arrays 
(ch3dif.f) 


Compute  constants 
(ch3div.f) 


Set  wind  stress 
(ch3dws.f) 


Is  flushing  model  included? 
(_FLUSHING_MODEL_) 


F  * 


^          Is  tidal  forcing         ~~^_ 
<C_                   read  from  file? 

-^(ITIDE  <  0)  ^ " 

--^             T 

Read  tidal  forcing  from  file 
(ch3dtd.f) 

F 

1 

r 

Details  in  Chart  2 

Main  time  iteration  loop 
(ch3dm2.f) 

f  END  J 

Figure  A.l  continue. 


Read  flushing  input  file 
(ch3dirjlu.f) 


Set  initial  concentration 
field  (ch3dics.f) 


266 


Chart  2:  main  time-iteration  loop  (ch3dm2.f) 


Called  from  ch3d.f :  Chart  1 


-lag  for  initiaT 

surface  elevation 

XISURF  =  3L 


Interpolate  the  initiatial 

surface  elevation 

(ch3dinitsurf.f) 


Set  depth  array 
(ch3ddp.f) 


Initializing  sediment  model 
Chart  3-1 


Initialize  sediment 

model 

(ch3dsedi.f) 


Initializing  water  quality  model: 
Chart  4-1 


initialize 

water  quality  model 

(ch3dwaini.f) 


Set  temperature 

at  river  boundary 

(ch3dtk.f) 


Initialize  river 
arrays(ch3dri.f) 


Calculate  3-D  velocity 

at  river  boundaries 

(ch3dri_3d.f) 


Is  precipiation 

^/evaporation  included?^ 

(IRAIN  *  0) 


Initialize  rainfall 

/evaporation  arrays 

(ch3drain.f) 


Figure  A.2  Flow  chart  for  the  driving  subroutine  for  the  time  stepping  of  the  solution. 


267 


Initialize  heavily  used 

transport  coefficients 

(ch3dcoef.f) 


Start  of  time  loop 


Time  varying  inflow 

river  discharge  routine 

(ch3drv.f) 


Calculate  water 
density&arroclinic 
pressure  in 
sigma  plane 
(ch3dde_i.f) 
(ch3ddej.f) 


ISIGZ=1 


CSD 


Calculate  water 

density&arroclinic 

pressure  in 

z  plane 
(ch3dde_z_i.f) 
(ch3dde_z_j.f) 


1s  vertical  eddy- 
coefficient  varying? 
(IEXP*01 


Computer  variable 

turbulent  eddy  coefficients 

(ch3ded.f) 


Figure  A. 2  continue. 


268 


Update  internal  and  external  variables 
(ch3dreset.f) 


Set  wind  field 
(ch3ddwt.f) 


T 

oes  rainfall  vary  witff-^ 

Update  the  time 
varying  rain  (ch3drt.f) 

time?(IRAIN  >  0 \^ 

^F 

Is  real  tide  being 
used?(ITIDE<0) 


Read  tide  data  and 
interpolate  (ch3dti.f) 


Ts  harmonic  tide  beinc] 
used?  (ITIDE>0L 


Update  old  depth 


Tidal  forcing  calculated 

from  harmonic  input  file 

(ch3dtd.f) 


Compute  Ul"*1  and  S*  for2-D  model 
(ch2dxy_i.f) 


Compute  Vl~'  and  S"-1  for  2-D  model 
(ch2dxy_j.f) 


Update  new  depth  arrays 
(ch3ddp.f) 


Figure  A. 2  continue. 


269 


Update  velocities 

at  river  boundaries 

(Ch3dri_3d.f) 


Compute  velocities  in  x/y/z  directions 
(ch3dxyz_i.f) 
(ch3dxyzj.f) 
(ch3dxyz_k.f) 


Compute  inertia&diffusion 

for  3-D  model  in  x/y  directions 

(ch3ddi_i.f) 

(ch3ddij.f) 


Compute  inertia&diffusion 

for  2-D  model  in  x/y  directions 

(ch2ddi_i.f) 

(ch2ddi_j.f) 


Is  salinity 

being  simulated? 

(IT>ISALT  & 

ISALT#0) 


Salinity  transport 
(ch3dsa.f) 


-Is  temperature^ 

being  simulated? 

(IT>ITEMP  & 

ITEMP*0) 


Temperature 
transport  (ch3dte.f) 


Suspended  sediment 
transport  (ch3dsed.f) 


Sediment  model: 
Chart  3-2 


Figure  A. 2  continue. 


(2D_2) 


270 


Water  quality  model 
Chart  4-2 


Water  quality  transport 

and  chemical  reactions 

(ch3dwq.f) 


Conservative 

species  transport 

(ch3dcs.f) 


Conservative 

species  transport 

for  2-D  model 

(ch3dcs.f) 


Check  conservation 
(ch3dcheckcons.f) 


Output  Hydrodynamics&salinity 
model  results(ch3dot.f) 


Output  sediment&wave 

model  results 

(ch3dot_sed.f) 


Run  "poor  mans  RPC 
via  tecplot  (ch3dpmrpc.O 


End  of  time  loop 


Return  to  chart 


Figure  A. 2  continue. 


271 


Chart  3-1:  Sediment  initialization  and  fetch  (ch3dsedi.f) 


Called  from  ch3dm2.f:  Chart  2 


Read  sediment  input  file  and  wave 

induced  bottom  shear  stress 

(ch3dirs.f) 


Initialization 
(ch3disd.f) 


Compute  fetch 
(ch3dft.f) 


.Return  to  ch3dm2.f:  Chart  2) 


Compute  wave  induced  bottom 

shear  stress 

(ch3dbs.f) 


Figure  A.3  Flow  chart  for  the  initializing  sediment  transport  model 


272 


Chart  3-2:  Sediment  model  (ch3dsed.f) 


Called  from  ch3dm2.f:  Chart  2 


T 


Compute  wave  induced  bottom 
shear  stress 

(ch3dbs.f) 


Calculate  settling,  deposition  velocities  and  erosion  rate  for 
fine  and  coarse  sediment 


Fine  group  advection,  diffusion 
(ch3dsd1  .f) 


Coarse  group  advection,  Diffusion 
(ch3dsd2.f) 


Calculate  sediment  bottom  change 


Return  to  ch3dm2.f:  Chart  2) 


Figure  A.4  Flow  chart  for  main  sediment  transport  model 


273 


Chart  4-1  :  Initializing  Water  Quality  Model  (ch3dwqini.f) 


Called  from  ch3dm2.f  :chart  2 


Initialize  WQ  variables 


Read  water  quality  input  file 
(ch3dir_nut.f) 


Read  dissolved  oxygen  data  for  several 

stations  (fort. 310)  and  time  and  space 

interpolate  (ch3drdoxy.f) 


F    •* 


Read  incident  light  data  (fort.307) 
and  time  interpolate  (ch3drlight.f) 


Read  temperature  data 

for  each  segment(fort.306) 

and  time  interpolate 

(ch3drtemp.f) 


Read  algae  growth  rate  for  each  segment 
(fort.309)  and  time  interpolate  (ch3dragrm.f) 


Read  color  data  for  each  segment 
(fort.308)  and  time  interpolate  (ch3drcolor.f) 


Figure  A.5  Flow  chart  for  initializing  water  quality  model 


274 


Sgad  initial  WQ  at  aerobic  sediment  layer. 
(INIT_WATER  t  0) 


Read  light  information:Table  for  absorption  rate 
according  to  wave  length  from  fort.321 


Read  all  initial  WQ  parameters 
at  computational  grid  points  (fort.99) 


Read  initial  WQ  for  water  column(fort.303) 
and  spatial  interpolate  (ch3dinterp.f) 


Read  initial  WQ  for  aerobic  sediment  column 
(fort.301)  and  spatial  interpolate  (ch3dinterp.f) 


Read  initial  WQ  for  anaerobic  sediment  layer 
(fort.302)  and  spatial  interpolate  (ch3dinterp.f) 


Set  initial  WQ  area  for  conservation  check 
and  initialize  assimilation  terms 


Read  river  boundary  data  from  fort.304  or  fort.305 
time  interpolate  (ch3drrivwq.f) 


F   K 


Return  to  ch3dm2.f :  chart  2 


Figure  A.5  Continue. 


275 


Chart  4-2  :  Water  Quality  Model  (ch3dwq.f) 


Called  from  ch3dm2.f  :chart  2 


i 


Set  up  assimilation  terms 
and  Reset  WQ  variables 


Read  river  boundary  WQ  data  (fort.305) 
and  time  interpolate  (ch3drrivwq.f) 


Calculate  horizontal  &  vertical  advection,  horizontal  diffusion 
in  water  column  and  specify  tidal  and  river  boundary  (ch3dtrp.f) 


Calculate  solar  radiation  angle 
(solar_angle.f) 


Compute  incident  light 
(ch3dlig.f) 


Read  incident  light  data  (fort.307) 
and  time  interpolate  (ch3drlight.f) 


Read  color  data  for  each  segment 
(fort.308)  and  time  interpolate  (ch3drcolor.f) 


Read  dissolved  oxygen  data  for  several 

stations  (fort.310)  and  time  and  space 

interpolate  (ch3drdoxy.f) 


Read  temperature  datafor  each  segment 
(fort.306)and  time  interpolate(ch3drtemp.f) 


Read  algae  growth  rate  for  each  segment 
(fort.309)  and  time  interpolate  (ch3dragrm.f) 


Figure  A.6  Flow  chart  for  main  water  quality  model. 


276 


Setup  Z-  vertical  grid 

include  water  and  sediment  column 

(gridld.f) 


Compute  temperature 

and  light  functions 

(temlit.f) 


Temperature  and  light  coefficient 
:  Chart  5 


Compute  vertical  diffusion  and  chemical  reaction 

for  plankton  species  in  water  column 

(ch3dalg.f) 


Prepare  variables  and  convert  to  z-direction 

1  dimensional  variables  in  water  and  sediment  column 

for  WQ  simulations  (preset.f) 


Compute  vertical  diffusion  and  chemical  reaction 

for  nitrogen  species  in  water  and  sediment  column 

(ch3dnitro.f) 


Compute  vertical  diffusion  and  chemical  reaction 

for  phosphorous  species  in  water  and  sediment  column 

(ch3dphosp.f) 


Compute  DO  and  CBOD 
(ch3doxy.f) 


Post  processing  of  WQ  variables 
(postset.f) 


Check  conservation  fro  WQ  species 
(ch3cons_wq.f) 


Output  water  quality  model  results 
(ch3ot_nut.f) 


Return  to  ch3dm2.f :  chart  2 


Figure  A. 6  continue. 


Chart  5:  Sets  up  temperature  and  light  functions  for  water  quality  model 
and  finds  light  intensity  in  a  particular  water  column  (temlit.f) 


277 


Set  up  temperature  functions 
For  water  quality 


Set  coefficient  for  algal  growth 
based  on  light  intensity 


Return  to  ch3dwq.f:  Chart  4-1 


Figure  A.7  Flow  chart  for  computing  the  temperature  and  light  attenuation  functions  for 
water  quality  model. 


APPENDIX  B 
DIMENSIONLESS  EQUATIONS  IN  CURVILINEAR  BOUNDARY-FITTED  AND 

SIGMA  GRID 

Non-dimensionalization  of  governing  equations  make  it  easy  to  compare  the  relative 

importance  of  various  terms  in  the  equations.     The  governing  equations  are  non- 

dimensionalized  using  the  following  reference  scales:  Xr  and  Zr  are  the  reference  lengths 

in  the  vertical  and  horizontal  directions;  [/,  is  the  reference  velocity;  p ,  pr  and 
Ap  =  p-  p0  are  the  reference  density,  mean  density  and  density  gradient  in  a  stratified 
flow;   AHr   and   AVr    are  the  reference  eddy  viscosities  in  the  horizontal  and  vertical 

directions;  Dw  and  DVr  are  the  reference  eddy  diffusivities  in  the  horizontal  and  vertical 
directions.  The  dimensionless  variables  can  be  written  as  (Sheng,  1983) 

(x\y\z*)  =  (x,y,zXr/Zr)/X, 

(u\v\w*)  =  (u,v,wXr/Zr)/Ur 
m  =coXrIUr 
t*=tf  (B.l) 

(f>;)  =  (r;,<)/(Poizr[/r)  =  (r;>;j/rr 
C  =  gCi(furxr)=cisr 

A\r  =  A/  '  AVr 

D'H=DH/  DHr 
D*v=Dv/DVr 

278 


279 

These  dimensionless  variables  can  be  combined  to  yield  the  following  dimensionless 
parameters 

Rossby  number  :  R  = 


Froude  Number  :  F  = 


°  4^ 
u. 


r  M 

F 

Densimetric  Froude  Number  :  J7     = !_  /g  ?^ 


Vertical  and  Horizontal  Ekman  Number  :  Ev  =  —^- ,  EH  =  —&- 


ft  ft 


A  A 

Vertical  and  Horizontal  Schmidt  Number  :  Srv  =—^-,Sru  =  -Jt- 

Dvr  DHr 


(3  =  ^ 


A, 

2  v2~  — 


KFrJ 


In  three-dimensional  modeling,  complex  bottom  topographies  can  be  better 
represented  with  the  application  of  o-stretching  (Sheng,  1983).  This  transformation  allows 
the  same  vertical  resolution  in  the  shallow  coastal  areas  and  the  deeper  navigation  channels. 
The  vertical  coordinate,  z,  is  transformed  into  a  new  coordinate,  o,  by  (Phillips,  1957). 

^_       z-£(x,y,t) 

h(x,y)  +  C(x,y,t)  (B3) 

where  h  is  the  water  depth  and  Q  is  the  surface  elevation. 

In  this  new  vertical  coordinate,  the  vertical  velocity  is  calculated  by  the  following 
equation. 


280 


DC 
w  =  H  co  +  (\  +  a)—?-  +  a 

Dt 


dh       dh 
u  —  +  v — 
dx       dy 


) 


(B.4) 


u  dz  da  .    , 

where  w  =  —   in  the  z-plane,  CO  = in  the  o-plane 

dt  dt 

Using  non-orthogonal  boundary-fitted  horizontal  grid,  it  is  possible  to  better  represent 
the  circulation  and  transport  processes  in  estuarine  systems  with  complex  shoreline 
geometries.  Using  the  elliptic  grid  generation  technique  developed  by  Thompson  (1982)  and 
Thompson  et  al.  (1985),  a  non-orthogonal  boundary-fitted  grid  can  be  generated  in  the 
horizontal  dimension.  To  solve  for  flow  in  a  boundary-fitted  grid,  it  is  necessary  to 
transform  the  governing  equations  from  original  coordinates  (x,  y)  to  the  transformed 
coordinates  ($,  r\).  The  spatial  coordinate  system  in  the  computational  plane  (£,  r|)  is 
dimensionless  while  the  coordinates  in  the  physical  plane  (x,  y)  have  dimensions  of  length. 
During  the  transformations,  the  velocities  are  transformed  into  contra-variant  velocities. 

In  the  boundary-fitted,  curvilinear,  o-stretched,  non-dimensional  coordinate  system, 
the  continuity  and  momentum  equations  are 


dt 


^(^»+4(^» 


do 


(B.5) 


281 


1  dHu 
H~dT 


( 


s  d£  *  drj 


S~       ,    8 

rll  +       , V 


^V  60 


^ 


■Vn 


_a_ 
a 


(  y{  4^HUU  +  y„  \[8~0HUV)  +  T-  (  Vf  V^# "V  +  ^  yfg^Hw} 


Ro 


-So" 


«£( 


a//i 


(B.6) 


wv 


ao- 


ap  +  12a^ 
a<f       a;/. 


Jcr  + 


^fr^flfr^H 


+ 


— 7- —  A^ —         +EHAH  (Horizontal     Diffusion    of     u) 
H   da\      da j 


i  a//v 

ff  ar 


d#   *  a^ 


+ 


.21 


\ 


U  +      , V 

"n/So 


+AL 


So# 


a_ 
a_ 

dHvw 


So 


f: 


«t 


da  + 


+ 


eu  a 


H2  do- 


da  j 

21  ^P  22  ^/° 

dv\ 
\- —        +EHAH  (Horizontal     Diffusion     of     v) 
da  ) 


2ldH        ,2dH 


[l#° 


+  crp 


(B.7) 


is  the  determinant  of  the  matric  tensor,  gu ,  which  is  defined  as 


(B.8) 


§"- 


x:  + 


i  +  ye 


xsx,  +  y*yn 


_xnx{  +  y,yz  xv  +  y, 


Sn     812 

621         <?22 


(B.9) 


whose  inverse  is 


282 


g'J  = 


4  +  yi  -(x^n  +  y{yn) 

-<XnXe+yvye)  x]+y] 


11  12 

8       8 
8       8 


(B.10) 


As  shown  in  Sheng  (1986),  the  contravarient  components  («')  and  physical 
components  (u(I))  of  the  velocity  vector  in  the  non-Cartesian  system  are  locally  parallel  or 
orthogonal  to  the  grid  lines,  while  the  covarient  components  («,)  are  generally  not  parallel 
or  orthogonal  to  the  local  grid  lines.  The  relationship  between  the  physical  velocity  and 
contravarient  velocity  is  given  by  (Sheng,  1986) 


u(i) 


with  summation  on  i. 


The  salinity  and  temperature  transport  equations  can  be  written  as 


dHS 

dt 


HSCV  da 


3  fDv^ 
vda 


K 


dHwS 
da 


& 


VioW 


±(4J-0HuS)  +  ±(^0HvS) 


+ 


+ 


Sch\I8o 


d  (  dS 


2  3S" 
) 


d    (    l—rr    2.  dS  f—  „    22  dS  ^ 


(B.ll) 


(B.12) 


and 


283 


BHT 

dt 


HSCV  da 


3  (dX 

v  da 


K 


H 


+ 


go 


dHwT 
da 


ToHuT)  +  —  (JT0HvT) 


*Z 


Scny[g~oldZ 


drj 

8oHgn—+Jg0Hg12  — 

d£  drj 


scH^ldrl 


i — _.  21 dT      / —     „dT^ 
d£  drj 


(B.13) 


The  sediment  transport  equation  can  be  written  as 


dHc: 


dt       H  ■  Scv  da 


H 


I  (D  V 


V 


-/?, 


y 


d(Hco-wJci 
da 


+ 


Ev 

ScHyJgo 

ScH\jgo 

f 


K 


dc 


goHg"-j:  +  JgoHgi--± 


V 

d f 


^ 


,  3c, 


3?7 


drj 

„dc" 


\ 


g0Hg2^  +  Jg~0Hg 


(B.14) 


where  c,  represents  cohesive  (i=l)  and  non-cohesive  (i=2)  sediment  concentrations  and  wVi 
is  settling  velocity  for  sediment  group  i. 


APPENDIX  C 
COMPARISON  OF  WATER  QUALITY  MODELS-WASP,  CE-QUAL-ICM,  and 

CH3D-WQ3D 


The  CH3D  water  quality  model  (CH3D-WQ3D)  was  compared  with  existing  water 
quality  models,  specially  Water  Quality  Analysis  ans  Simulation  Program  (WASP),  the 
integrated  Compartment  water  quality  model  developed  by  the  US  Army  Corps  (CE-QUAL- 
ICM).  The  methods  of  coupling  with  hydrodynamic  and  sediment  transport  models,  the 
simulated  parameters,  the  assumptions,  the  chemical/biological  processes,  and  the  limitations 
of  each  model  are  discussed  and  compared  in  Table  C.l.  The  first  criterion  in  model 
comparison  is  its  ability  to  simulated  hydrodynamic,  sediment  transport,  and  water  quality 
with  the  efficient  coupling. 

The  WASP  box  modeling  framework  has  proven  to  be  an  excellent  water  quality 
model  for  riverine  systems,  where  the  steady  state  assumption  is  applicable.  In  addition,  this 
simple  box  model  can  be  successfully  used  to  perform  numerical  experiments  like  sensitivity 
test.  However,  in  marine  environments,  it  should  not  be  used  without  proper  linkage  with 
3-D  hydrodynamics  and  sediment  models,  because  tide  wind  and  baroclinic  forcing  underact 
in  an  unsteady  balance. 

As  pointed  out  by  Chen  and  Sheng  (1994),  the  loosely  coupled  models  such  as  CE- 
QUAL-ICM,  cannot  account  for  nutrient  release  by  sediments  in  episodic  events.  Because 
these  models  were  not  coupled  with  a  dynamic  model  for  sediment  transport,  they  could  not 


284 


285 
accurately  consider  sediment-process  effects  such  as  resuspension,  deposition,  flocculation, 

and  settling  on  nutrient  dynamics  in  estuaries.   Furthermore,  this  model  use  equilibrium 

partition  with  function  of  biomass  for  hydrolysis  process.  This  may  be  a  reasonable 

assumption  when  the  time  step  of  the  water  quality  simulation  is  large  compared  to  the  time 

it  takes  to  reach  equilibrium.  However,  sometimes  a  water  quality  model  may  use  a  small 

time  step  during  which  the  absorbed  and  the  dissolved  nutrient  may  not  achieve  equilibrium. 

In  this  case,  a  kinetics  model,  such  as  sorption/desorption  kinetic  process  in  CH3D-WQ3D, 

is  needed.   Analytical  chemists  repeatedly  found  that  complete  recovery  of  contaminants 

from  soil/sediments  frequently  requires  lengthy  extraction  periods,  abrasive  mixing,  and 

strong  solvents  (Witkowski  and  Jaffe,  1978).  These  observations  are  in  contradiction  with 

the  equilibrium  models  which  assume  that  sorption-desorption  reactions  are  accomplished 

instantaneously. 

The  light  attenuation  model  need  chlorophyll  a  and  suspended  sediment 
concentrations  to  calculate  light  attenuation  in  water  column.  The  water  quality  model  use 
light  intensity  for  limitation  of  algae  growth  rate  and  photosynthesis  processes.  Therefore, 
without  fully  coupling  these  models,  it  could  not  be  accomplished  to  communicate  each 
others  with  both  direction. 

The  chemical/biological  processes  in  all  three  models  are  basically  similar  and 
compatible  with  each  others.  Most  of  them  are  the  adjustable  coefficients.  Thus,  the  ability 
to  simulate  water  quality  depends  on  modeler  experience,  and  necessity  of  model  complexity. 
Also  the  data  availability  is  the  key  to  ensure  reliability  and  accuracy  of  the  results.  The  CE- 
QUAL-ICM  include  much  more  simulating  species  than  those  for  the  other  model.  Without 
measured  data  for  those  species,  these  complex  nutrient  cycles  could  create  more  uncertainty 


286 
than  simple  nutrient  species  model.  CE-QUAL-ICM  use  sediment  flux  model  developed  by 

Ditoro  and  Fitzpatric  (1990)  for  sediment  oxygen  demand.  Although,  applying  of  sediment 
flux  model  provides  rational  predictions  of  sediment  response  to  environmental  alterations, 
it  increases  information  requirements  and  computation  time  compare  to  employment  of  user- 
specified  fluxes  (Cerco  and  Cole,  1995).  They  point  out  that  employment  of  user-specified 
fluxes  show  the  better  calibration  of  the  water  quality  model  than  employment  of  the 
sediment  flux  model. 

Overall,  with  the  similarity  and  compatibility  of  chemical/biological  processes  which 
is  depend  on  data  availability,  coupling  with  hydrodynamic  model,  sediment  transport  model, 
temperature  model,  and  light  attenuation  model  is  the  key  to  ensure  reliability  and  accuracy 
of  model.  Specially,  the  Charlotte  Harbor  estuarine  system  which  has  a  strong  linkage 
between  hydrodynamics  and  water  quality  dynamics  suggests  the  importance  of  coupling 
with  hydrodynamics  and  sediment  transport.  Therefore,  CH3D-WQ3D  is  better  to  simulate 
water  quality  in  Charlotte  Harbor  estuarine  system  than  the  other  water  quality  models. 


287 


Table  C.l  Comparison  of  water  quality  models 


model 

WASP5.X 

CE_QUAL_ICMvl.O 

CH3D-WQ3D 

Hydrodynamic 
model 

Read  into  the 
model  as  input 
parameters 

read  into  the  model 
as  input  parameters 

fully  coupled  with 
circulation,  salinity 
and  temperature 
transport  model 

Sediment 
transport  model 

Read  into  the 
model  as  input 
parameters 

loosely  coupled  with 
simple  model 

fully  coupled  with 
sediment  dynamic 
model 

Light  attenuation 
model 

loosely  coupled 
with  empirical 
formula 

loosely  coupled  with 
empirical  formula 

fully  coupled  with 
physics-based  light 
attenuation  model 

Nitrogen  cycle 

NH3,  N03,  ON 

NH4,  N03,  SON, 
LPON,  RPON 

NH4,  N03,  SON, 
PON,  PIN,  NH3 

phosphorous 
cycle 

SRP,  OP 

SRP,  SOP,  LPOP, 
RPOP 

SRP,  SOP,  POP, 
PIP 

carbon  cycle 

Using  CBOD 

SOC,  LPOC,  RPOC 

Using  CBOD 

silica  cycle 

No 

Available  Silica, 
Particulate  Biogenic 
silica 

No 

metal  cycle 

No 

Iron  &  Manganese 

No 

phytoplankton 

One  species 

Three  species 

One  species 

zooplankton 

Using  zooplankton 
grazing  rate 

Using  phytoplankton 
predation  rate 

Yes 

Metabolism 

Respiration  and 

non-predator 

mortality 

One  parameter 

Respiration  and 

non-predator 

mortality 

re-aeration 

Function  of  current 
velocity  and  depth 

Constant 

Function  of  current 
velocity,  depth,  and 
wind  speed 

sediment  oxygen 
demand 

Empirical  formula 

Sediment  flux  model 

Empirical  formula 

Variation  of 
Ammonification 

Function  of 
temperature 

Function  of  algae 
biomass 

Function  of 
temperature 

Variation  of 
mineralization 

Function  of 
temperature 

Function  of  algae 
biomass 

Function  of 
temperature 

288 


Hydrolysis 

No  particulate 

Equlibrium  partition 

Sorption/desorption 

species 

with  function  of 
algae  biomass 

kinetic 

phosphorous 

Constant 

Variable  ratio 

Constant 

/carbon  ratio 

described  by 

empirical 

approximation 

APPENDIX  D 
NUMERICAL  SOLUTION  TECHNIQUE  FOR  WATER  QUALITY  PROCESSES 


In  the  finite  difference  solution  of  the  water  quality  model,  the  advection  and 
horizontal  diffusion  terms  are  treated  explicitly,  whereas  the  vertical  diffusion  and 
biogeochemical  transformations  are  treated  implicitly.  Fractional  step  methods,  which 
guarantee  numerical  stability  and  prevent  negative  concentrations,  are  applied  in  the 
numerical  solution  (Chen  and  Sheng,  1994).  The  horizontal  diffusion  and  horizontal  and 
vertical  advection  terms  are  solved  first.  The  numerical  solutions  proceeds  with  the 
calculation  of  vertical  diffusion,  and  then  biogeochemical  transformation  reactions.  Finally, 
the  sorption/desorption  reaction  terms  are  solved.  Equation  (D.l)  shows  a  schematic  of  the 
numerical  solution  algorithm  method  used  in  this  study. 


At 

Nn2-Nn{ 


-  [Horiz.  Advection  +  Vertical  Advection  +  Horiz.  Diffusion]" 

■  =  [Vertical  Diffusion]"2  +  [Q]"2  (D.l) 


At 

=  [Sorption]"    +  [desorption] 


Nn+l-N"2     r„        .     in+1     r,  .     in+1 


At 

By  solving  the  sorption/desorption  terms  separately  from  other  terms,  it  is  possible 
to  treat  these  terms  implicitly.  To  illustrate  this,  the  difference  equations  for  the 
sorption/desorption  reactions  are  examined: 


289 


290 
d  ^   d  =HdxN?-dx-Px.c-N?1) 

N>Hl  _  Nn2  (D.2) 

'  ^    '  =<dxN?-dx.px-c-N?x) 

where  Nd  and  Np  are  concentrations  of  dissolved  and  particulate  nutrients  such 
as,NH4  and  PIN,  SON  and  PON,  SRP  and  PIP,  and  SOP  and  POP;  dx  is  desorption  rate  for 
a  specific  nutrient  species;  andp^  is  partition  coefficient  for  a  specific  nutrient  species. 

In  solving  the  second  step  of  the  fractional  step  method  for  organic/inorganic 
phosphorous  and  nitrogen  species,  the  difference  equations  for  the  entire  water  column  and 
sediment  column  are  solved  simultaneously.  The  kinetic  processes  for  particulate  nutrient 
and  plankton  species  need  to  include  a  settling  process,  which  accounts  for  the  limited 
vertical  motion  of  these  species. 


For  dissolved  species 


For  particulate  species 


dN 


d    _ 


dt       dz 


D. 


dN 


\ 


<i 


dz 


+  Q  (D.3) 


dND      a 


f 


dz  J 


+  Q  (DA) 


dt       dz 

The  exchanges  of  particulate  nutrients  at  the  water-sediment  interface  are  determined 
by  the  sediment  resuspension  and  deposition  fluxes  while  the  exchange  of  dissolved  nutrient 
species  between  the  water  and  sediment  column  are  automatically  included  in  the  numerical 
solutions,  with  diffusion  term.  In  order  to  ensure  that  the  water  quality  model  is  consistent 
with  the  sediment  model,  in  terms  of  bottom  exchanges,  the  erosion  and  deposition  rates 
calculated  in  the  sediment  model  can  be  imported  into  the  governing  equations  for 
particulate  nutrients.  Therefore,  the  particulate  nutrient  concentrations  are  solved  with  the 
unit  of  percentage  (jig/jig).  Let  p  be  any  particulate  species  (Np)  mass  per  unit  mass  of 
sediments,  then, 


N=p-c 


291 
(D.5) 


where  c  is  the  suspended  sediment  concentration.  Substitute  Np  with  pc  into  second  step  of 
fractional  step  method: 


dpc  _  d 
dt       dz 


wspc  +  Dv 


dpc 

~dz~ 


+  Q 


dc       dp      d 
p  —  +  c— =  — 
dt        dt      dz 


wc  +  DV  — 

vdz 


dz 


v   dz 


+  Q 


(D.6) 


(D.7) 


Since  the  vertical  one-dimensional  sediment  equations  are: 


dc       d  (  dc^ 

—  =  —   wtc  +  Dv  — 
dt      dzy  '  dz 


(D.8) 


Equation  (D.7)  becomes: 

dt 


wc  +  D  — 

'dz 


dz     dz 


dA 

v    dz 


+  Q 


(D.9) 


This  equation  allows  us  to  import  the  erosion  and  deposition  rates  calculated  in  the 
sediment  model  into  the  particulate  nutrient  species.  Since  the  boundary  condition  for  the 
sediment  model  at  the  water-sediment  interface  is: 


wsc  +  Dv-^-  =  D-E 
dz 


Equation  (D.9)  at  the  water-sediment  interface  becomes: 


(D.10) 


dt  dz     dz 


dz 


+  Q 


(DM) 


where  D  and  E  are  deposition  and  erosion  rate,  which  are  calculated  at  sediment  model. 

Due  to  the  difference  in  the  partial  pressures  of  oxygen  and  carbon  and  because  of 
other  physical  and  biochemical  factors,  the  transformation  processes  in  the  water  column  and 
in  the  sediment  column  are  different.  There  are  no  phytoplankton  and  zooplankton  species 


292 


in  the  sediment  column.  When  these  plankton  species  die,  they  are  treated  as  particulate 
organic  species.  In  the  sediment  column,  there  exist  two  distinctive  layers:  an  aerobic  layer 
and  an  anaerobic  layer,  within  which  the  transformation  processes  are  not  the  same.  For 
example,  in  the  anaerobic  layer,  there  is  no  nitrification  process  due  to  a  lack  of  oxygen, 
while  in  the  aerobic  layer,  there  is  no  denitrification  process  due  to  the  availability  of  oxygen. 
The  vertically,  one  dimensional  z-grid  is  shown  in  Figure  D.l. 


a> 
o 

Q. 

-o 


KMG 


KMG-1 


CO 

o 

3 


o 
o 
3 


KMS+2 


KMS+1 


KMS 


KMR+1 


KMR 


V 


KMG=KM+KMS+KMR 


Water  Column 
KMS=KMR+KMO 


7^ 

o 


-V 


Aerobic  Sediment  Column 
KMR 


J3 


Anaerobic  Sediment  Column 


Figure  D.l  The  vertical  one-dimensional  z-grid 

Since  the  horizontal  transport  is  generally  very  weak  in  the  sediment  column,  the 
mass  flows  of  nutrient  species  are  mainly  vertical,  and  the  governing  equations  are: 


For  dissolved  species  in  a  sediment  layer 


dt       dz 


0M 


dz  > 


+  Q        (D.12) 


293 


For  particulate  species  in  a  sediment  layer 


dNn 

wrN+M p- 

c    P  dz 


+  Q    (D.13) 


dt       dz 

where  Nd  is  dissolved  nutrient  species  such  as  NH4,  SON,  SRP,  SOP  concentrations  (per  unit 
volume  of  porewater),  wc  is  the  consolidation  velocity  of  sediments,  M  is  the  molecular 
diffusivity,  and  0is  the  porosity. 


APPENDIX  E 
NUTRIENT  DYNAMICS 

Nutrients  are  essential  elements  for  life  processes  of  aquatic  organisms.  Nutrients 
of  concern  include  carbon,  nitrogen,  phosphorous,  silica  and  sulfur.  Among  these  nutrients, 
the  first  three  elements  are  utilized  most  heavily  by  zooplankton  and  phytoplankton.  Since 
carbon  is  usually  available  in  excess,  nitrogen  and  phosphorous  are  the  major  nutrients 
regulating  the  ecological  balance  in  an  estuarine  system.  Nutrients  are  important  in  water 
quality  modeling  for  several  reasons.  For  example,  nutrient  dynamics  are  critical 
components  of  eutrophication  models  since  nutrient  availability  is  usually  the  main  factor 
controlling  algae  bloom.  Algae  growth  is  typically  limited  by  either  phosphorous  or 
nitrogen.  (Bowie  et  al.,  1980) 

Nutrient  inputs  to  estuarine  systems  are  related  to  point  and  non-point  sources  from 
land,  atmospheric  deposition,  and  fixation.  Additionally,  internal  loadings  such  as  from 
resuspended  sediments  containing  inorganic  and  organic  forms  are  also  important.  The 
specification  and  quantification  of  each  of  these  contributions  are  the  first  steps  towards  the 
determination  of  nitrogen  and  phosphorous  budgets  in  an  estuarine  system. 

Nutrient  cycles  are  highly  dependent  on  the  hydrodynamics  and  sediment  dynamics 
of  the  estuarine  system.  Resuspension  events,  combined  with  desorption  processes,  can 
significantly  change  the  input  and  budget  of  nitrogen  and  phosphorous  in  the  system.  On  the 
other  hand,  deposition  and  sorption  may  contribute  to  major  losses  of  nitrogen  and 


294 


295 
phosphorous  from  the  water  column.   The  hydrodynamics  not  only  derive  the  sediment 

processes,  but  also  affect  the  sorption/desorption  reactions,  through  turbulent  mixing. 
E.l  Nitrogen  Cycle 

Nitrogen  can  be  classified  into  two  groups:  dissolved  nitrogen  and  particular  nitrogen. 
The  criterion  of  this  division  is  established  in  the  laboratory  using  filtering  technique.  The 
dissolved  nitrogen  include  ammonia  nitrogen  (NH3),  dissolved  ammonia  nitrogen  (NH4), 
nitrite  and  nitrate  nitrogen  (N03),  and  dissolved  organic  nitrogen  (SON).  Particulate 
nitrogen  includes  particulate  inorganic  nitrogen  (PIN),  and  particulate  organic  nitrogen 
(PON).  Phytoplankton  nitrogen  (PhyN)  and  zooplankton  nitrogen  (ZooN)  related  biomass 
to  nitrogen  concentration  through  a  fixed  stoichiometric  ratio:  nitrogen-to-carbon  ratio  (ANC). 

The  model  nitrogen  cycle  (Figure  E.l)  includes  the  following  processes. 

1)  Ammonification  of  organic  nitrogen 

2)  Nitrification  of  ammonium 

3)  Volatilization  of  ammonia 

4)  Denitrification  of  nitrate 

5)  Uptake  of  ammonia  and  nitrate  by  phytoplankton 

6)  Conversion  of  phytoplankton  nitrogen  to  zooplankton  nitrogen  by  grazing 

7)  Excretion  and  mortality  by  phytoplankton  and  zooplankton 

8)  Settling  for  particulate  nitrogen 

9)  Sorption/desorption  reactions 


296 


tVolatili 


NH3 


Volatilization 

NH3-(NH3),J 

Instalbility 


Excretion  (Kzx'ZOON) 


PIN 


Kal — SS. —  NH4 

Hal  +  pH 


Sorption/Desorption 

d„(PIN-p„*C-NH4 


"   " 


Zooplankton 


Mortality 

(1-KPDN)*K   -ZOON 


NH4 


Nitrogen 


Diffusion*     Erosion/ 


Amonification 


Mortality 

KPDhTK   'ZOON 


SON 


Mortality 

(1-KPDN)*Kas'PHYN 


Uptake  Pm"ua*PHYN 


Excretion  K./PHYN 
Nitrification 


Sorption/Desorptio 


,(PON-p    'C'SO 


on 


PON 


Mortality 

KPDN'K   -PHYN 


Phytoplankton 


Kim ^?— AW4 

Hnit+DO 

♦  Diffusion 


N03 


Uptake 


(1-PJ'u-PHYN 


f.|iIhsS:-;:<      4 


•■     .    .<■     "'       .:, 


PIN 


Sorption/Desorption 


°,„(PIN-P„-C-NH    * 
Instalbility 


NH4 


Amonification 


SON 


1 


Water  Column 

fcrosioii/    ^Diffusion 


iorption/Desorptio 


„(PON-p   "C 


>rption 

•SON! 


NH3 


Kal  ^ AW  4r 

Hal  +  pH 


PON 


Nitrification 

Knn - AW  4 

<  <        Hnil  +  DO 


N03 


♦  Diffusa 


"t"""*"""""" 


N03 


Denitrification 


Kiln       "n°3       A-03 
Hno3+  DO 


N2 


Aerobic  Layer 


Anaerobic  Layer 


Figure  E.l  Nitrogen  Cycle 

Ammonification  is  the  biological  process  of  formation  of  ammonium  from  soluble 
organic  nitrogen.  It  is  the  first  step  of  nitrogen  mineralization,  in  which  organic  nitrogen  is 
converted  to  the  more  mobile,  inorganic  state.  The  rate  of  ammonification  is  expressed  as 
a  first  order  reaction  (Rao  et  al,  1984) : 


0, 


ammonification  **  ONM   '  "  " ^ 


(E.l) 


where  Kom  is  the  rate  constant  of  ammonification  which  is  a  function  of  water  temperature, 
pH,  and  the  C/N  ratio  of  the  residue  (Reddy  and  Patrick,  1984). 

The  second  step  of  mineralization  of  organic  nitrogen  is  nitrification,  which  is  an 


oxidation  of  ammonia  to  nitrate(  NO^  )  directly  or  to  nitrite  ( NO:  )  and  then  to  nitrate: 


297 

nh: + 1 .5a  ->  no; +ih++ h.o 

(E.2) 

Nitrification  requires  oxygen  as  the  electron  acceptor.  Therefore,  nitrification  is  a 
strictly  aerobic  process,  occurring  only  in  the  water  column  and  in  the  aerobic  layer  of  the 
sediment  column.  This  process  is  related  to  a  sink  of  dissolved  oxygen  in  the  system  as 
shown  in  eq  4.11.  The  kinetics  of  nitrification  are  modeled  as  function  of  available 
ammonia,  dissolved  oxygen,  and  temperature. 

^nitrification  ~  ~  ^  NN  ~7l  nTTi"  4  (E-3) 

Mnit+DO 

where  KNN  is  nitrification  rate  which  is  a  function  of  temperature;  and  Hni,  is  the  half- 
saturation  constant  for  the  bacteria  growth. 

The  dissolved  form  of  ammonium  in  water  is  generally  not  stable  and  can  exist  in  its 
gaseous  form,  or  ammonia  nitrogen.  Because  the  ammonia  concentration  in  the  atmosphere 
is  very  low,  ammonia  in  the  water  column  can  escape  to  the  air.  This  is  a  volatilization 
process  of  ammonia.  The  volatilization  of  ammonia  is  a  sink  for  nitrogen  in  an  aquatic 
system.  A  first  order  rate  equation  can  be  used  to  describe  the  kinetics  of  the  ammonia 
volatilization  process  (Chen  and  Sheng,  1994): 

&***.  =  Kvol  [KNH3  -  NH°°*  ]  (E.4) 

where  KV0L  is  rate  constant  of  volatilization  which  is  a  function  of  temperature.  It  can 

be  derived  from  the  so-called  two-film  model  (Jorgensen,  1983);  hv  is  henry's  constant;  and  NH"'m 

is  the  ammonia  concentration  in  the  air. 

Denitrification  is  defined  as  the  biogeochemical  transformation  of  nitrate  nitrogen  to 
gaseous  end  products  such  as  molecular  nitrogen  or  nitrous  oxide  (Reddy  and  Patrick,  1 984). 


298 
While  nitrification  occurs  in  the  water  column  and  aerobic  layer  of  the  sediment  column, 

denitrification  occurs  only  in  the  anaerobic  layer  of  the  sediment  column.  The  denitrification 
process  can  be  described  by  the  standard  Michaelis-Menten  equation  (Bowman  and  Focht, 
1974). 

jj 

U denitrification  =~^dn~  "  ~ ~  "03  (E.5) 

Hno3  +  DO 

where  Kdn  is  the  denitrification  rate,  which  is  a  temperature  function;  and  Hn^  is  the  half- 
saturation  constant  for  denitrification. 

In  the  nitrogen  species,  sorption  processes  refer  to  conversion  from  a  soluble  to  a 
solid  phase  of  inorganic  (NH4  to  PIP)  and  organic  (SON  to  PON)  species,  while  desorption 
reactions  describe  the  inverse  process.  Sorption/desorption  processes,  combined  with 
resuspension  events  can  significantly  alter  the  nitrogen  cycle  in  the  system.  The  kinetics  of 
sorption/desorption  reactions  are  dependent  on  nitrogen  species  characteristics,  sediment 
properties,  pH,  temperature,  and  dissolved  oxygen  concentration  (Simon,  1989)  The  most 
commonly  used  mathematical  representation  of  sorption/desorption  processes  is  the  linear, 
reversible,  isotherm  (Berkheiser  et  al.,  1980;  Reddy  et  al.,  1988): 

JtNad--Dr-Nud+Sr-Ns  (E.6) 

where  Dr  is  the  desorption  rate  constant  which  is  a  temperature  function;  Sr  is  the  sorption 
rate  constant;  Nad  is  the  adsorbed  nitrogen  concentration  such  as  particulate  inorganic  and 
organic  nitrogen;  N,  is  the  dissolved  nitrogen  concentration  such  as  dissolved  ammonium 
nitrogen  and  dissolved  organic  nitrogen. 

The  ratio  between  the  desorption  and  sorption  rates  gives  the  partition  coefficient 
dissolved  and  particulate  forms,  because  dNad/dt  =  0  at  equilibrium. 


299 


Sr  _  <d  _ 


o  =  Pc  (E.7) 


Dr      N 
where  A/^and  AT"  are  the  adsorbed  and  dissolved  nitrogen,  respectively,  at  the  equilibrium 

condition;  and  /?,.  is  the  partition  coefficient.     Therefore,  the  kinetic  equation  for 
sorption/desorption  reaction  is: 

JtNad=-Dr-(Nad-pc-Ns)  (E.8) 

Inorganic  nitrogen  is  incorporated  by  phytoplankton  during  growth  and  release  as 
ammonium  and  organic  nitrogen  through  respiration  and  non-predatory  mortality.  The 
phytoplankton  nitrogen  can  be  converted  to  zooplankton  nitrogen  by  grazing  process.  The 
kinetic  processes  for  particulate  nitrogen  species  need  to  include  settling  process,  which 
accounts  for  the  limited  vertical  motion  of  particulate  nitrogen.  For  this  species,  it  is 
reasonable  to  assume  the  same  settling  velocity  of  the  suspended  sediment  particles. 

The  mass  balance  equations  for  nitrogen  state  variables  are  written  by  combining 
nitrogen  transformation  processes. 

Ammonia  nitrogen  (NH3) 

include  ammonia  conversion  and  volatilization  processes 


For  water  column 

d  ._„       v  pH 


dt  Hal  +  pH 


hv-NH3-NHu3""]  (E.9) 


For  sediment  column: 


dt  Hal  +  pH 

where  Kal  is  the  ammonia  conversion  rate  constant  which  is  a  temperature  function;  Hal  is 

half-saturation  constant  for  ammonia  conversion. 
Dissolved  ammonium  nitrogen  (NH4) 


300 
include     phytoplankton     uptake     and     respiration,     zooplankton     respiration, 
ammonification,  nitrification,  ammonia  conversion,  and  sorption/desorption  reaction. 
For  water  column: 

-NH4=-[(Pn.Ma- Kax ) •  PhyC - Ku  ■  ZooC] ■  ANC  +  K0NM  ■  SON 

-Km — NH4-Kal P— NH4  (E-11) 

Hnil+DO         4       -   Hal+PH         4 

+dan(PIN-pan-c-NH4) 
For  sediment  column: 

-NH4=+KONM.SON-KNN.-^.NH4-Kar--^--.NH4 

m  Hni,+D0  Hal  +  pH  (E.12) 

+dan(PIN-Pan-c-NH4) 

where  dan  is  sorption/desorption  rate  of  NH4  from  sediment  particles;  pan  is  the  partition 
coefficient  between  NH4  and  PIN;  and  c  is  the  suspended  sediment  concentration. 

Nitrate  and  nitrite  nitrogen  (NQ3) 

include  nitrification,  denitrification  and  phytoplankton  uptake  processes. 

For  water  column: 

|-/V03  =  +KNN  ■       D°       ■  NH4  -  KDN ^ NO, 

dt  Hnil+DO  Hlw3+DO        3  (e.13) 

-ANCia-Pn>Ma'PhyC] 

For  sediment  column: 

Soluble  organic  nitrogen  (SON) 

include  ammonification  and  sorption/desorption  reaction. 

For  water  column: 

-SON  =  -KONM  -SON  +  don(PON-pm-C'SON)  (E.15) 

For  sediment  column: 


301 

^SON  =  -K0NM  ■  SON  +  d(m  -(PON-  Pon  -c-SON)  (E.16) 

where  d(m  is  sorption/desorption  rate  of  SON  from  sediment  particles;  andpon  is  the  partition 
coefficient  between  SON  and  PON 

Particulate  organic  nitrogen  (PON) 

include  mortality  of  phytoplankton  and  zooplankton,  settling,  and  a  sorption- 

desorption  reaction. 

For  water  column: 

-PON  =  ANC\Kas-PhyC  +  Kzs.ZooC]--wsp-PON 

-don(PON-Pon-c-SON) 
For  sediment  column: 

—  PON  = ws  •  PON  -  don  ( PON  -  pon  ■  c  ■  SON )  (E.  18) 

dt  dz 

where  wsp  is  settling  velocity  for  particulate  species,  which  is  same  with  that  of  suspended 

sediment  particles. 

Particulate  inorganic  nitrogen  (PIN) 

include  settling  and  a  sorption/desorption  reaction. 

For  both  water  and  sediment  columns: 

^PIN  =  -j-wsp-PIN-dan(PIN-pan-c-NH4)  (E.19) 

E.2  Phosphorous  Cycle 

Phosphorous  can  be  classified  into  two  groups:  dissolved  phosphorous  and  particulate 
phosphorous.  The  criterion  for  this  division  is  established  in  the  laboratory,  using  a  filtering 
technique.  The  dissolved  phosphorous  include  soluble  reactive  phosphorous  (SRP)  and 
dissolved  organic  phosphorous  (SOP).  Particulate  phosphorous  includes  particulate 
inorganic  phosphorous  (PIP),  and  particulate  organic  phosphorous  (POP).  Phytoplankton 


302 
phosphorous  (PhyP)  and  zooplankton  phosphorous  (ZooP)  related  biomass  to  phosphorous 

concentration,  through  a  fixed  stoichiometric  ratio:  phosphorous-to-carbon  ration  (APC). 

The  model  phosphorous  cycle  (Figure  E.2)  includes  the  following  processes: 

1)  Mineralization  of  organic  phosphorous 

2)  Uptake  of  soluble  reactive  phosphorous  by  phytoplankton 

3)  Conversion  of  phytoplankton  phosphorous  to  zooplankton  phosphorous  by  grazing 

4)  Excretion  and  mortality  by  phytoplankton  and  zooplankton 

5)  Settling  for  particulate  phosphorous 

6)  Sorption/desorption  reactions 


Excretion  (Kzx'ZOOP) 


PIP 


Sorption/Desorption       „_„ 

<upip-p -c-srpT     "HK 


Zooplankton 


Mortality 

(1-KPDP)*K   'ZOOP 


Mineralization 


Mortality 
KPDP'K   -ZOOP 


SOP 


Mortality 

(1-KPDPCK   'PHYP 


Uptake  P  -u  "PHYP 


Excretion  K. /PHYP 


Sorption/Desorption 

V(pop-p  -c-sopT 


POP 


Mortality 

KPDP'K   "PHYP 


Phytoplankton 


Phosphorous 


Diffusion  ♦     Erosion/ 


*  Diffusion 


Diffusion  * 


Water  Column 

Erosion/    4  Diffusion 


~ 


t  deposition 


PIP 


Sorption/Desorption 


"dJPIP-p   -C'SRPy 


_ 


deposition 


SRP 


Mineralization 


SOP 


Sorption/Desorption         „_„ 
VJPOP-p-C'SOrf        KUK 


♦  Diffusion 

T"     *" 


p.p         Sorption/Desorption 


♦  Diffusion  Diffusion^ 

Mineralization 


SRP 


Aerobic  Layer 
Diffusion^ 

"T ' 


SOP 


Sorption/Desorption 

a„(POP-P  -c-sopT 


POP 


opv  ^op 


Anaerobic  Layer 

Figure  E.2  Phosphorous  Cycle 

The  mineralization  process  is  mediated  by  bacteria,  which  transfers  dissolved  organic 
phosphorous  to  soluble  reactive  phosphorous,  through  the  uptake  of  SOP  and  excretion  of 


303 
SRP.  Since  bacteria  abundance  is  related  to  algae  biomass,  the  rate  of  organic  phosphorous 

mineralization  is  related  to  algae  biomass.  The  mineralization  of  SOP  is  a  relatively  fast 

process,  it  can  take  a  few  hours,  compared  to  the  mineralization  of  carbon  and  nitrogen, 

which  takes  place  in  a  few  days  (Golterman,  1973).  Mineralization  is  highest  when  algae  are 

strongly  phosphorous  limited  and  is  lowest  when  no  limitation  occurs.   Thompson  et  al. 

(1954)  found  that  the  mineralization  of  dissolved  organic  phosphorous  is  influenced  by  pH 

value.    An  increase  in  pH  causes  a  temporary  increase  in  the  rate  of  mineralization  of 

dissolved  organic  phosphorous.  Temperature  can  also  affect  the  speed  of  the  mineralization 

by  stimulating  the  mineralization  process  with  high  temperatures. 

The  mineralization  rate  of  SOP  is  usually  modeled  by  a  first  order  equation,  as 

follows  (Jorgensen,  1983) 

^mineralization  ~        "■  opm  '  ^U*  (E.20) 

where  Koptn  is  a  rate  constant  for  mineralization  of  SOP,  which  is  a  function  of  pH  and 
temperature. 

Soluble  reactive  phosphorous  is  incorporated  by  phytoplankton,  during  growth  and 
release,  as  soluble  reactive  phosphorous;  and  organic  phosphorous  through  respiration  and 
mortality.  The  phytoplankton  phosphorous  is  converted  to  zooplankton  phosphorous  by 
grazing  processes.  The  settling  and  sorption/desorption  processes  are  similar  to  the  nitrogen 
species. 

The  mass  balance  equations  for  phosphorous  state  variables  are  written  by  combining 
these  phosphorous  transformation  processes. 

Soluble  reactive  phosphorous  (SRP) 

include  mineralization,  uptake  by  phytoplankton,  mortality  of  zoo  and  phytoplankton, 


304 
and  sorption/desorption  reaction. 

For  water  column: 

T- SRP  =  KoP>„  ■  SOP  +  Apc  [-jua  ■  PhyC  +  K^  ■  PhyC  +  K.x  ■  ZooC] 
m  (E.21) 

+dip(PIP-pip-c-SRP) 

For  sediment  column: 

JtSRP  =  Kopm  ■  SOP  +  dip  (PIP -Pip-c- SRP)  (E.22) 

where  dip  is  sorption/desorption  rate  of  SRP  from  sediment  particles;  and/?,p  is  the  partition 
coefficient  between  SRP  and  PIP. 

Soluble  organic  phosphorous  (SOP) 

include  mineralization  and  sorption/desorption  reaction. 
For  both  water  and  sediment  columns: 

-SOP  =  -Kopm  -SOP-d^POP-p^c-SOP)  (E.23) 

where  dl)p  is  sorption/desorption  rate  of  SOP  from  sediment  particles;  andpop  is  the  partition 
coefficient  between  SOP  and  POP. 

Particulate  organic  phosphorous  (POP) 

include    respiration    of    zooplankton    and    phytoplankton,    settling,    and    a 
sorption/desorption  reaction. 


For  water  column: 

d  d 

—  POP  =  APC  ( Kux  ■  PhyC  +  K^  -ZooC)-  —  wsp-  POP 

-dop-(POP-Pop-c-SOP) 
For  sediment  column: 


(E.24) 


jtPOP  =  -  —  wspPOP-dop\POP-pop-c-SOP)  (E.25) 


305 
Particulate  inorganic  phosphorous  (PIP) 

include  settling  and  a  sorption/desorption  reaction 
For  both  water  and  sediment  columns: 

-PIP  =  -—wSp-PIP-dip-(PIP-Pip.cSRP)  (E.26) 


APPENDIX  F 
SEDIMENT  FLUX  MODEL 

DiToro  et  al.  (1990)  developed  a  model  of  the  SOD  process  in  a  mechanistic  fashion 

using  the  square-root  relationship  of  SOD  to  sediment  oxygen  carbon  content.  Using  similar 

analysis  as  applied  to  carbon,  they  also  evaluate  the  effect  of  nitrification  on  SOD.  In  this 

model,  carbon  and  nitrogen  diagenesis  are  assumed  to  occur  at  uniform  rates  in  a 

homogeneous  layer  of  the  sediment  of  constant  depth  (active  layer).  The  sediment  oxygen 

demand  and  sediment  fluxes  are  calculated  by  simple  first-order  decay  processes  of  the 

concentrations  of  particulate  organic  carbonaceous  material  and  of  particulate  organic 

nitrogenous  material  in  this  active  layer  as  follow: 

dC 

poc 


=  -kpocCpoc+Mc  (F.l) 

at 


dC 


,  K pon^ pon  ~ ln  N  \L  --J 

where  CP0C  is  a  concentration  of  POC  in  sediment  (g  02/m3) 
CP0N  is  a  concentration  of  PON  in  sediment  (g  02/m3) 
KP0C  is  a  decay  rate  of  POC  in  sediment  (g  02/m3) 
KPON  is  a  decay  rate  of  POC  in  sediment  (g  02/m3) 
MN  is  a  source  term  for  CP0C  (g/m3-day) 
Mc  is  a  source  term  for  CPON  (g/m3-day) 
Sediment  carbon  and  nitrogen  diagenesis  fluxes,  Jc  and  JN,  are  the  most  important 


306 


307 
parameters  in  the  equations  for  sediment  oxygen  demand  and  sediment  fluxes  as  defined  by 

h  =  ~kpocCpocH 

i      -b    r    h  F3) 

JN  ICpon(^pon" 

where  H  is  the  depth  of  the  active  layer  (m) 

Using  these  diagensis  fluxes,  DiToro  and  Pitzpatrick  developed  sediment  flux  model 
which  include  NSOD,  CSOD,  ammonia  and  nitrate  flux,  methane  flux,  and  sulfide  flux. 
F.l  Sediment  Flux  Equations  inCH3D-WQ3D 

The  reactive  portion  of  particulate  organic  carbon  and  particulate  organic  nitrogen 
in  the  sediment  are  presented  by  CBOD  and  NH4  in  CH3D-WQ3D  model  system.  The 
equation  (F.3)  can  be  rewritten  as 

Jc=-kgxy-Q™CBODstd.H 

(R4) 
JN=-KM<i,-°-NHAsed-H 

where  koxy  is  a  oxidation  rate  in  sediment  (g  02/m3) 
kni[  is  a  nitrification  rate  in  sediment  (g  02/m3) 

®l~y°  is  a  temperature  effect  of  oxidation 

@Jt'  is  a  temperature  effect  of  nitrification 

In  this  study,  sediment  fluxes  include  sediment  oxygen  demand,  benthic  dissolved 
methane  flux,  and  benthic  methane  gas  flux.  Sediment  oxygen  demand  (SOD)  can  be 
determined  for  the  set  of  equations  developed  by  DiToro  and  Fitzpatrick  (1993). 


Carbonaceous  sediment  oxygen  demand  (g/m2/d): 


CSOD  =  J2KDCSJC 


f 


1-sec/z 


KrO, 


SOD 

Note:  The  square  root  term  is  replaced  by  Jc  ifJc  <2KDC, 

Nitrogenous  sediment  oxygen  demand  (g/m2/d): 


(F.5) 


308 


/ 


NSOD  =  1.714  J> 


Total  sediment  oxygen  demand  (g/m2/d): 


1  -  sec  h 


V 


KNQ2 
SOD 


(F.6) 


SOD  =  CSOD  +  NSOD  (F.7) 

where  02  is  a  concentration  of  dissolved  oxygen  in  overlying  water  column  (g  02/m3). 

kD  is  a  dissolved  methane  diffusion  mass  transfer  coefficient  (g  02/m3) 

kc  is  a  reaction  velocity  for  methane  oxidation  (g  02/m3) 

kN  is  a  reaction  velocity  for  ammonia  oxidation  (g  02/m3) 

C,  is  a  methane  solubility  (g  02/m3) 

The  magnitude  of  the  fluxes  at  the  sediment  water  interface  of  aqueous  methane  and 
gaseous  methane  are  predicted  to  be 


•* CHA(aq)         V  *>      S     C 


sec/z 


Kc02 
SOD 


"'CH4(g)~''C        V  O      5     C 


(F.8) 


(F.9) 


The  original  DO  kinetic  equation  in  CH3D-WQ3D  is  modified  to  incorporate  the 
sediment  exchange  predictions  of  the  DiToro  model,  given  by  equation  (F.10). 


dDO 
dt 


-Oxydation  -  Nitrification  +  reaeration  +  photosynthesis  -  respiration 


SOD        JcH4(aq)        ^  "* CH4(g) 


(F.10) 


H  H  J,u'      H 

where  Gfrac  is  a  fraction  of  gaseous  methane  produced  in  the  sediment. 

His  a  water  depth. 

The  first  five  source  and  sink  terms  in  the  equation  (F.10)  are  the  same  with  the 
original  DO  kinetic  equation,  while  the  last  three  terms  represent  oxygen  demand  due  to 
sediment  exchange  processes,  where  SOD  and  aqueous  and  gaseous  methane  flux  rates. 


APPENDIX  G 
MODEL  PERFORMANCE  TEST  WITH  PARALLEL  CH3D-IMS 

To  reduce  the  computational  time  necessary  to  simulate  multiple-year  seasonal 
response  of  the  Charlotte  Harbor  estuarine  system  with  a  serial  CH3D  model,  a  parallel 
CH3D  model  was  developed  and  validated.  The  parallel  approach  is  applied  to  the  parallel 
CH3D  model  via  parallel  constructs  added  to  the  serial  CH3D  model  (Davis  and  Sheng, 
2000;  Sheng  et  al.,  2003).  These  parallel  constructs  are  implemented  by  adding  additional 
macros  to  the  original  serial  source  code.  By  defaults,  the  parallel  CH3D  model  uses 
OpenMP  constructs  although  due  to  the  flexible  nature  of  the  implementation  process,  either 
Sun  Microsystems-style  or  Clay-style  constructs  can  be  used.  The  parallel  source  code 
closely  resembles  the  serial  source  code  and  maintains  perfect  compatibility  with  the  serial 
code.  In  other  words,  the  parallel  CH3D  model  can  be  executed  using  the  same  input  files 
as  the  serial  CH3D  model  and  parallel  CH3D  output  format  is  identical  to  the  output  format 
of  the  serial  CH3D  model.  To  determine  how  well  the  parallel  CH3D  model  performs  in 
Charlotte  Harbor  estuarine  system,  the  serial  and  parallel  CH3D  simulations  performed  for 
the  parallel  model  validation. 
G.l  Validation  of  Parallel  Model 

To  validate  the  parallel  CH3D  model,  simulations  are  performed  using  both  the  serial 
and  parallel  CH3D  model.  CH3D  model  output  from  all  of  the  parallel  simulations  are  then 
compared  to  their  respective  serial  simulations  to  fully  validate  the  parallel  CH3D  model. 


309 


310 
The  computational  grid  for  Charlotte  Harbor  estuarine  system  contains  92  x  129  horizontal 

cells  (Figure  5 . 1 )  and  8  vertical  layers,  with  a  total  1 1 648  grid  cells  which  include  5367  water 

cells  and  6281  land  cells.  Grid  spacing  varies  from  40  to  2876  meters  (average  598  meters). 

Using  this  grid,  the  3-D  hydrodynamic  and  water  quality  simulation  simulate  the  circulation, 

sediment  transport,  and  water  quality  dynamics  of  the  Charlotte  Harbor  estuarine  system 

from  May  23  to  June  22, 1996.  The  initial  condition  for  simulation  is  provided  by  a  30-day 

spin  up  simulation  (April  23  to  May  23)  previously  performed  during  the  dry  season  with  all 

forcing  mechanisms  (tides,  river  discharges,  wind)  to  allow  water  level,  velocity  and  salinity 

field  to  reach  dynamic  steady-state  throughout  the  computational  domain.  A  more  detailed 

description  of  the  simulation  can  be  found  in  Chapter  6. 

Simulated  time  series  and  field  parameters  at  the  end  of  30-day  simulation  of  each 
parallel  simulation  are  compared  to  the  all  corresponding  simulated  parameters  from  the 
serial  CH3D  simulation  which  include  water  level,  current  velocity,  temperature,  salinity, 
sediment  concentration,  and  all  water  quality  species.  For  all  comparisons,  the  simulated 
time  series  and  field  parameters  obtained  from  the  serial  CH3d  model  are  identical  to  the 
simulated  parameters  obtained  from  the  parallel  CH3D  model.  Thus  the  parallel  CH3D 
model  is  completely  validated. 
G.2  CPU  Times  of  Parallel  Routines 

The  1 -month  simulation  is  simulated  using  both  serial  and  parallel  CH3D  models  on 
the  SGI  orgin  system  witch  is  Silicon  Graphyics  Origin  3400,  400  MHz  MIPS  R 12000  (IP 
35).  This  platform  has  16  processors  and  8  GB  main  memory  size.  The  parallel  model  is 
executed  using  from  1  to  4  processors  on  this  system.  Table  F.l  report  the  CPU  times  per 
iteration  for  the  parallel  CH3D  model  executed  on  this  platform.  This  timing  results  show 


311 
that  the  time  necessary  for  CH3D  simulation  can  be  significantly  reduced  using  multi- 
processor computers  coupled  with  parallel  techniques. 

Table  G.l  CPU  time  for  the  parallel,  shared  memory,  CH3D  procedures  on  SGI  origin 
platform.  Time  shown  are  per  time  step  iteration  of  the  model  using  computational  grid 
(92x  1 29)  and  are  given  in  seconds,  n  is  the  number  of  processors  used  and  speed  up  is  shown 
in  parenthesis. 


Serial 


n=1 


n=2 


n=3 


n=4 


Main  WQ 

1.221(1.00) 

1.287(0.95) 

0.680(1.80) 

0.469(2.60) 

0.370(3.30) 

Main  Sediment 

0.126(1.00) 

0.125(1.01) 

0.078(1.62) 

0.060(2.10) 

0.052(2.41) 

Turbulence 

0.100(1.00) 

0.107(0.93) 

0.059(1.70) 

0.045(2.21) 

0.038(2.61) 

Sediment  transport 

0.090(1.00) 

0.089(1.01) 

0.047(1 .94) 

0.032(2.85) 

0.025(3.62) 

Dimensionalize 

0.075(1 .00) 

0.083(0.90) 

0.047(1.62) 

0.033(2.28) 

0.026(2.89) 

Baroclinic  (J) 

0.071(1.00) 

0.075(0.95) 

0.042(1 .70) 

0.038(1.87) 

0.036(1.97) 

Baroclinic  (I) 

0.070(1.00) 

0.067(1.05) 

0.038(1.86) 

0.025(2.75) 

0.020(3.49) 

Salinity 

0.055(1.00) 

0.059(0.95) 

0.030(1.85) 

0.020(2.51) 

0.016(3.43) 

N.L/Diffusion  (J) 

0.055(1.00) 

0.055(1.00) 

0.033(1.65) 

0.030(1 .84) 

0.029(1.87) 

N.L/Diffusion  (I) 

0.042(1 .00) 

0.043(0.98) 

0.025(1.65) 

0.018(2.34) 

0.015(2.82) 

Layer  Vel.  (v) 

0.040(1.00) 

0.041(0.97) 

0.027(1.48) 

0.021(1.86) 

0.019(2.12) 

Layer  Vel.  (u) 

0.028(1 .00) 

0.029(0.97) 

0.016(1.73) 

0.011(2.60) 

0.009(3.28) 

Layer  Vel.  (w) 

0.011(1.00) 

0.011(0.99) 

0.006(1.70) 

0.004(2.37) 

0.003(2.85) 

Interpolation 

0.011(1.00) 

0.011(1.00) 

0.006(1.79) 

0.005(2.62) 

0.004(3.52) 

integrate  Vel.  (U) 

0.008(1 .00) 

0.008(1.00) 

0.006(1.34) 

0.004(1.69) 

0.004(1.87) 

integrate  Vel.  (V) 

0.006(1.00) 

0.006(0.97) 

0.004(1.35) 

0.005(1 .32) 

0.005(1 .26) 

Bottom  sheer  stress 

0.001(1.00) 

0.001(1.01) 

0.001(1.89) 

0.000(2.76) 

0.000(3.60) 

Wave  H/T 

0.001(1.00) 

0.001(1.03) 

0.000(1.95) 

0.000(2.93) 

0.000(3.87) 

All  Parallel  Routine 

1.973(1.00) 

2.060(1.00) 

1.152(1.71) 

0.843(2.34) 

0.701(2.82) 

Total  Routine 

2.022(1 .00) 

2.107(1.00) 

1.204(1.68) 

0.895(2.26) 

0.753(2.69) 

G.3  Parallel  Speedup 

The  parallel  speed  up  for  the  Charlotte  Harbor  1 -month  simulation  described  earlier 
on  the  SGI  Origin  platform  is  shown  in  Figure  G.l.  As  the  lines  get  close  to  the  theoretical 
maximum,  the  parallel  model  is  performing  better.  The  speedups  of  the  individual  parallel 
procedures  are  shown  as  values  inside  parenthesis  in  the  previous  table  G.  1  and  also  illustrate 


312 
how  computationally  intense  routines  have  higher  speedups.  Applying  parallel  method  for 
CH3D  model  achieved  a  1.71x  and  2.82x  speedups  using  2  processors  and  4  processors, 
respectively. 


2  3 

Numbers  of  Processors 

Figure  G.l  Parallel  speedup  gained  in  performing  the  simulation  on  the  SGI  Origin 
platform 


APPENDIX  H 
TIDAL  BENCH  MARKS  FOR  CHARLOTTE  HARBOR 

Table  H.l  Tidal  datum  referred  to  Mean  Low  Low  Water  (MLLW),  in  meter. 

Station       Station  Latitude     Longitude     MHHW     MHW     MLW      NAVD 

Number     Name 


8725853 

Venice 

27  04.3 

82  27.3 

0.671 

0.589 

0.113 

0.496 

8725791 

Peace 
River 

26  59.3 

81  59.6 

0.615 

0.537 

0.119 

0.501 

8725781 

Shell 
Creek 

26  58.8 

81  57.6 

0.668 

0.570 

0.135 

0.472 

8725541 

Bokeelia 

26  42.4 

82  09.8 

0.526 

0.481 

0.072 

0.495 

8725520 

Fort  Myers 

26  38.8 

81  52.3 

0.401 

0.335 

0.191 

0.318 

8725391 

Sanibel 

26  29.3 

82  00.8 

0.689 

0.614 

0.145 

0.570 

8725110 

Naples 

26  07.8 

81  48.7 

0.874 

0.797 

0.184 

0.696 

313 


REFERENCES 

Ambrose,  R.  B.  Jr.,  Wool,  T.A.,  Martin,  J.L.,  Connolly,  J.P.,  Schanz,  R.W.  (1994)  WASP5, 
A  hydrodynamic  and  water  quality  model-model  theory,  user's  manual,  and  programmer's 
guide.  U.S. EPA  Environmental  Research  Laboratory.  Athens,  Georgia. 

American  Public  Health  Association(APHA)  (1985).  Standard  methods  for  the  examination 
of  water  and  wastewater.  APHA.  Washington,  D.C. 

Banks,  R.B.  (1975).  Some  features  of  wind  action  on  shallow  lakes,  ASCE,  Journal  of 
Environmental  Engineering  Division,  Vol.  101,  No.  EE5,  p8 13-827 

Banks,  R.B.  and  Herrera,  F.F.  (1977).  Effect  of  wind  and  rain  on  surface  reaeration,  ASCE, 
Journal  of  Environmental  Engineering  Division,  Vol.  103,  No.  EE3,  p489-504 

Baker,  B.  (1990).  Draft  Caloosahatchee  water  quality  based  effluent  limitations 
documentation  (Lee  County).  Florida  Department  of  Environment  Regulation.  Water  Quality 
Tech.  Ser.  2.  No.  121:  18. 

Berkheiser,  V.E.,  Street,  J.J.,  Rao,  P.S.C.,  and  Yuan,  T.L.  (1980)  Partitioning  of  inorganic 
orthophosphate  in  soil-water  systems.  CRC  Critical  Review  in  Environmental  Control. 

Berliand,  M.E.,  and  Berliand,  T.G.  (1952).  Determining  the  net  longwave  radiation  of  the 
Earth  with  the  consideration  of  the  effect  of  cloudiness,  Isv.  Acad.  Nauk.,  SSSR  Ser.  Geofis, 
No.  1. 

Boler,  R.N.,  Malloy,  R.C.,  and  Lesnett,  E.M.  (1991)  Surface  water  quality  monitoring  by  the 
Environmental  Protection  Commission  of  Hillsborough  County.  In  (S.F.  Treat  and  Clark 
P.A.,  eds.)  Proceedings  of  Tampa  Bay  Area  Scientific  Information  Symposium  2,  1991. 
Tampa,  FL. 

Bowie,  G.L.,  Mills,  W.B.,  Porcella,  D.B.,  Campbell,  C.L.,  Pagenkopf,  J.R.,  Rupp,  G.L., 
Johnson,  K.M.,  Chan,  P.W.H.,  Gherini,  S.A.  and  Chamberlin,  C.E.  (1985).  Rates,  constants, 
and  kinetics  formulations  in  surface  water  quality  modeling.  In:  (Editors),  U.S. 
Environmental  Protection  Agency,  Office  of  Research  and  Development,  Athens,  GA 

Bowie,  G.L.,  Chen,  C.W.  and  Dykstra,  D.H.  (1980)  Lake  Ontario  ecological  modeling,  phase 
HI.  Tectra-Tech,  Inc.,  Lafayete,  CA. 


314 


315 

Bowman,  R.A.  and  Focht,  D.D.  (1974).  The  influence  of  glucose  and  nitrate  concentrations 
upon  denitrification  rates  in  sandy  soils.  Soil  Biol.  Biochem.  6:  297-301. 

Bricaud,  A.,  Morel,  A.,  and  Prieur,  L.  (1981).  Absorption  by  dissolved  organic  matter  of  the 
sea  (yellow  substance)  in  the  uv  and  visible  domains.  Limnology  and  Oceanography, 
28(5):43-53. 

Budyko,  M.  I.  (1974).  Climate  and  Life.  Academic  Press  Inc.,  New  York,  508pp. 

Camp,  Dresser,  &  Mckee  (1998).  The  Study  of  Seasonal  and  Spatial  Patterns  of  Hypoxia  in 
the  Upper  Charlotte  Harbor.  Prepared  for  S  WFWMD,  Prepared  by  Camp,  Dresser,  &  Mckee 

Canale,  R.P.,  DePalma,  L.M.,  and  Vogal,  A.H.  (1976).  A  plankton-based  food  web  model 
for  Lake  Michigan,  in  (R.P.  Canale,  ed.)  Modeling  Biochemical  Processes  in  Aquatic 
Ecosystem,  Ann  Arbor  Science  Publishers  Inc.,  P.O.  Box  1452,  Ann  Arbor,  Michigan  48 160. 

Cerco,  D.F.,  and  Cole,  T.M.  (1994).  Three-dimensional  eutrophication  model  of  Chesapeake 
Bay.  Technical  Report  El-94-4,  U.S.  Army  Engineer  Waterways  Experiment  Station, 
Vicksburg,  MS. 

Cerco,  D.F.,  and  Cole,  T.M.  (1995).  User  Guide  to  the  CE-QUAL-ICM  three-dimensional 
eutrophication  model,  Technical  Report  EL-95-15,  U.S.  Army  Corps  of  Engineers, 
Washington,  DC. 

Chamberlain,  R.H.  and  Doering,  P.H.  (1998).  Freshwater  inflow  to  the  Caloosahatchee 
Estuary  and  the  resource-based  method  for  evaluation.  Proceedings  of  the  Charlotte  Harbor 
public  conference  and  technical  symposium,  Technical  Report  No.  98-02:  81-90. 

Chapra,  S.C.  (1997).  Surface  water-quality  modeling.  The  McGRAW-HILL  companies, 
INC.,  New  York. 

Chen,  C.W.,  andOrlob,  G.T.  (1975).  Ecologic  simulation  of  aquatic  environments.  Systems 
Analysis  and  Simulation  in  Ecology,  Vol.  3,  B.C.  Pattern  (ed.).  Academic  Press,  New  York: 
475-588. 

Chen,  X.  and  Sheng,  Y.P.(1994).  Effects  of  hydrodynamics  and  sediment  transport  processes 
on  nutrient  dynamics  in  shallow  lakes  and  estuaries.  Ph.D.  dissertation.  University  of  Florida, 
Gainesville,  Florida 

Christian,  D.J.  and  Sheng,  Y.P.(2003).  Modeling  the  Effects  of  Hydrodynamics,  Suspended 
Sediments,  and  Water  Quality  on  Light  Attenuation  in  Indian  River  Lagoon,  Florida.  PhD 
dissertation,  University  of  Florida.,  Gainesville,  Florida. 

Clark,  N.  E.,  Eber  R.  M.  Renner  J.  A.,  and  Saur  J.  F.  T.  (1974)  Heat  exchange  between  ocean 
and  atmosphere  in  the  eastern  North  Pacific  for  196-71.  NOAA  Tech.  Rep.  NMFS  SSRF- 
682,  U.S.Dept.  Commerce,  Washington. ,D.C. 


316 

Davis,  J.R.  and  Sheng,  Y.P.  (2000).  High  performance  estuarine  and  coastal  environmental 
modeling:  The  CH3D  example.  In  Estuarine  and  Coastal  Modeling,  Vol.  6,  p470-486. 
American  Society  of  Civil  Engineers. 

Davis,  J.R.  and  Sheng,  Y.P.  (2001).  High  Performance  Modeling  of  Circulation  and 
Transport  in  the  Indian  River  Lagoon,  Florida.  PhD  dissertation,  University  of  Florida., 
Gainesville,  Florida. 

Day,  J.W.  Jr.,  Hall,  C.A.S,  Kemp,  K.M.,  and  Arancibia,  A.J.  (1989).  Estuarine  Ecology. 
John  Wiley  &  Sons.  New  York. 

DeCosmo,  J.,  Katsaros,  K.B.,  Smith,  S.D.,  Anderson,  R.J.,  Oost,  W.A.,  Bumke,  K.,  and 
Chadwick,  H.M.  (1996).  Air-sea  exchange  of  water  vapor  and  sensible  heat:  The  Humidity 
Exchange  Over  the  Sea  (HEXOS)  results.  J.  Geophys.  Res.-Oceans,  101:  12,001-12,016. 

Degrove,  B.  (1981).  Caloosahatchee  River  wasterload  allocation  documentation.  Florida 
Dept.  Of  Env.  Reg.,  Water  Quality  Tech  ser.  2.  No.  52:  17. 

Degrove,  B.  and  Nearhoof,  F.  (1987).  Water  quality  assessment  for  Caloosahatchee  River. 
Florida  Dept.  Of  Env.  Reg.  Water  Quality  Tech.  Ser.  3.  No.  19. 

DiToro,  D.M.,  Paquin,  P.R.,  Subburamu,K.,  and  Gruber,  D.A.  (1990).  Sediment  oxygen 
demand  modehMethan  and  ammonia  oxidation.  Journal  of  Evironmental  Engineering, 
116(5),  p945-986. 

DiToro,  D.M.  and  Fitzpatrick,  J.J.  (1993).  Chesapeaka  bay  sediment  flux  model.  U.S.  Army 
Corps  of  Engineers,  Waterways  Experiment  Station,  Tech.  Report  EL-93-2. 

Dixon,  L.K.,  and  Gary,  J.K.  (1999).  Cause  of  light  attenuation  with  respect  to  seagrasses  in 
upper  and  Lower  Charlotte  Harbor.  Marine  Laboratory  Technical  Report  No.  650.  Prepared 
for  Southwest  Florida  Water  Management  District. 

Doering,  P.H.  (1996).  Temporal  variability  of  water  quality  in  the  St.  Lucie  Estuary,  South 
Florida.  Water  Resource  Bulletin  32:1293-1306. 

Echternacht,  K.L.  (1975).  A  study  of  the  precipitation  regimes  of  the  Kissimmee  River-Lake 
Okeechobee  watershed.  Florida  Department  of  Environmental  Regulation  Technical  Service 
(3)1.  Tallahassee,  Florida. 

Eckart,  C.  (1958).  Properties  of  water.  Part  H  The  equation  of  state  of  water  and  sea  water 
at  low  temperatures  and  pressures.  American  Journal  of  Science,  256(4):  224-240. 

Edwards,  R.E.,  Lung,  W.,  Montagna,  P. A.,  and  Windom,  H.L.  (2000).  Final  review  report. 
Caloosahatchee  Minimim  Flow  Peer  Review  Panel,  September  27-29,  2000.  Report  to  the 
south  Florida  Water  Management  District,  West  Palm  Beach,  Florida. 


317 

Environmental  Quality  Laboratory,  Inc.  (1979).  Hydrobiological  monitoring  report  to  the 
Southwest  Florida  Water  Management  District,  Port  Charlotte,  Florida. 

Environmental  Quality  Laboratory,  Inc.  (1987).  Hydrobiological  monitoring  program,  data 
report  for  the  period  from  March  1986  through  Feburaury  1987  covering  the  lower  Peace 
River  abd  Charlotte  Harbor.  Port  Charlotte,  Florida,  Environmental  Quality  Laboratory,  Inc.: 
92. 

Eppley,  R.W.  (1972).  Temperature  and  phytoplankton  growth  in  the  sea.  Fish.  Bull., 
70:1063-1085. 

Estevez,  E.  D.  (1986).  Infaunal  macro  invertebrates  of  the  Charlotte  Harbor  estuarine  system 
and  surrounding  inshore  waters,  Florida..  U.S.  Geological  Survey  Water-Resources 
Investigations  Report  85-4260:  116. 

Estevez,  E.D.  (1998).  The  story  of  the  greater  Charlotte  Harbor  Watershed,  Charlotte  Harbor 
Estuarine  Program,  North  Fort  Myers,  Florida. 

Fraser,  T.H.  (1986).  Long-term  water-quality  characteristics  of  Charlotte  Harbor,  Florida. 
U.S.  Geological  Survey  Water-Resources  Investigations  Report  86-4180:  43. 

Fraser,  T.H.  and  Wilcox,  W.H.  (1981).  Enrichment  of  a  subtropical  estuary  with  nitrogen, 
phosphorous,  and  silica,  in  Neilson,  B.J.,  and  Cronin,  L.E.,  eds.,  Estuarine  and  Nutrients, 
Humana  Press,  New  Jersey:  481-498. 

Friehe,  C.  A.  and  Schmitt,  K.  B.  (1976).  Parameterization  of  air-sea  interface  fluxes  of 
sensible  heat  and  moisture  by  the  bulk  aerodynamic  formulas.  Journal  of  Physical 
Oceanography,  6,  801-809. 

Gallegos,  C.L.  (1993).  Determination  of  optical  water  quality  requirements  in  the  Indian 
River  near  Ft.  Pierce,  FL,  with  emphasis  on  the  impact  of  colored  water  discharges. 
Technical  report,  South  Florida  Water  Management  District. 

Gallegos,  C.L.  and  Correl,  D.L.  (1990).  Modeling  spectral  diffuse  attenuation,  absorption, 
and  scattering  coefficients  in  a  turbid  estuary.  Limnology  and  Oceanography,  35(7):  1484- 
1502. 

Gallegos,  C.L.  and  Kenworthy,  W.J.  (1996).  Seagrass  depth  limits  in  the  Indian  River 
Lagoon  (Florida,  USA).  Estuarine,  Coastal,  and  Shelf  Science,  42:267-288. 

Garret,  J.R.  (1977).  Review  of  drag  coefficients  over  oceans  and  continents.  Monthly 
Weather  Review  105:915-929. 

Golterman,  H.L.  (1973).  Vertical  Movement  of  Phosphorous  in  Freshwater,  in  (E.J.  Griffith, 
A.  Beeton,  J.M  Spencer,  and  D.T.  Mitchell,  eds.)  Environmental  Phosphorous  Handbook, 
John  Wiley  &  Sons.,  New  York. 


318 

Goodwin,  C.R.  (1996).  Simulation  of  tidal-flow,  circulation,  and  flushing  of  the  Charlotte 
Harbor  estuarine  system,  Florida.  U.S.  Geological  Survey  Water-Resources  Investigations 
Report  93-4153. 

Goodwin,  C.  R.  and  Michaelis,  D.M.  (1976).  Tides  in  Tampa  Bay,  Florida:  June  1971  to 
December  1973.  U.S.  Geological  Survey  Open-File  Report  FL-75004:  338. 

Hammet,  K.  M.  (1990).  Land  use,  water  use,  streamflow  characteristics,  and  water-quality 
characteristics  of  the  Charlotte  Harbor  inflow  area,  Florida.  U.S.  Geological  Survey  Water- 
Supply  Paper  2350-A:  64. 

Hammet,  K.  M.  (1992).  Physical  processes,  salinity  characteristics,  and  potential  salinity 
changes  due  to  freshwater  withdrawals  in  the  tidal  Myakka  River,  Florida.  U.S.  Geological 
Survey  Water-Resources  Investigations  Report  90-4054:20 

Heyl,  M.G.  (1997).  Hypoxia  in  upper  Charlotte  Harbor.  Proceedings  of  the  Charlotte  Harbor 
public  conference  and  technical  symposium,  Technical  Report  No.  98-02:  219-228. 

Hwang,  T.C.  (1966).  A  Sedimentologic  Study  of  Charlotte  Harbor,  Southwestern  Florida. 
Master's  thesis,  Florida  State  University,  Tallahassee,  Florida. 

Jaworski,  N.A.  (1981).  Source  of  Nutrients  and  the  scale  of  eutrophication  problems  in 
estuaries,  in  Neilson,  N.J.,  Jersey,  Humana  Press:  p83-110. 

Jerlov,  N.G.  (1976).  Marine  Optics.  Elsevier  Scientific  Publishing  Company,  New  York. 

Jorgensen,  S.E.  (1976).  A  eutrophication  model  for  a  lake.  Ecological  Modeling  2: 147-165. 

Jorgensen,  S.E.  (1983).  Modeling  the  ecological  processes.  In:  (G.T.  Orlob,  ed.) 
Mathematical  Modeling  of  Water  Quality:  Streams,  Lakes,  and  Reservoirs.  John  Wiley  and 
Sons.  New  York.  Pp  116-149. 

Jorgensen,  S.E.,  Sorensen,  B.H.,  and  Nielsen,  S.N.  (editors)  (1996).  Handbook  of 
Environmental  and  Ecological  Modeling.  CRC  Press,  Inc. 

Jorgensen,  S.E.,  and  Gromiec,  M.E.  (editors)  (1989).  Mathematical  sub-models  in  water 
quality  systems.  Development  in  Environmental  Modeling,  14.  Elsevier  Science  Publishers. 

Ketchem,  B.H.  (1967).  Phytoplankton  nutrients  in  estuaries,  in  Lauff,  G.  H.,  ed.,  Estuaries. 
National  Oceanic  and  Atmospheric  Administration 

Kirk,  J.T.O.  (1984).  Dependence  of  relationship  between  inherent  and  apparent  optical 
properties  of  water  on  solar  altitude.  Limnology  and  Oceanography,  29(2):350-356. 


319 

Lam,  D.D.L.,  Schertzer,  W.M.,  and  Fraser,  A.S.  (1983).  Simulation  of  Lake  Erie  water 
quality  responses  to  loading  and  weather  variations,  I.W.D.  Scientific  Series,  134, 
Environment  Canada,  Ottawa:299. 

McEwan,  J.,  Gabric,  A.J.  &  Bell,  P.R.F.  (1998).  Water  quality  and  phytoplankton  dynamics 
in  Moreton  Bay,  Southeastern  Queensland,  n.  Mathematical  modeling.  Marine  and 
Freshwater  Research,  Vol.  40,  No.  3,  p215-225. 

McPherson,  B.  F.  and  Miller,  R.L.  (1990).  Nutrient  distribution  and  variability  in  the 
Charlotte  Harbor  estuarine  system,  Florida.  Water  Resource  Bulletin  26:  67-80. 

McPherson,  B.  F.,  Miller,  R.  L.  and  Stoker,  Y.  E.  (1996).  Physical,  Chemical,  and  Biological 
Characteristics  of  the  Charlotte  Harbor  Basin  and  Estuarine  System  in  Southwestern  Florida- 
A  Summary  of  the  1982-89  U.S.  Geological  Survey  Charlotte  Harbor  Assessment  and  Other 
Studies.  U.S.  Geological  Survey  Water-Supply  Paper  2486. 

Metha,  A.J.  (1986).  Characterization  of  cohesive  sediment  properties  and  transport  processes 
in  estuaries.  In:A.J.  Metha  (ed.),  Estuarine  Cohesive  Sediment  Dynamics.  Springer,  New 
York:290-325. 

Mehta,  A.J.  (1989).  On  estuarine  cohesive  sediment  suspension  behavior.  Journal  of 
Geophysical  Research  94(C  10):  14303-14314. 

Mellor,  G.  L.,  (1996),  Introduction  to  Physical  Oceanography,  Springer,  Verlag  New  York. 

Miller  R.  L.  and  McPherson  B.  J.,  (1995).  Modeling  Photosynthetically  Active  Radiation  in 
Water  of  Tampa  Bay,  Florida,  with  Emphasis  on  the  Geometry  of  Incident  Irradiance. 
Estuarine,  Coastal  and  Shelf  Science,  40,  359-377  . 

Morel,  F.  (1983).  Principles  of  aquatic  chemistry.  John  Wiley  and  Sons,  New  York,  NY:  150. 

Morel,  A.,  and  Gentili,  B.  (1991).  Diffuse  reflectance  of  oceanic  waters:  Its  dependence  on 
sun  angle  as  influenced  by  the  molecular  scattering  contribution.  Applied  Optics  30:4427- 
2238. 

Morrison,  G.  (1997).  Proposed  trophic  state  goals  and  nitrogen  management  objectives  for 
the  tidal  reaches  of  the  Peace  and  Myakka  rivers.  Draft.  Prepared  by  Southwest  Florida  Water 
Management  District,  Surface  Water  Improvement  and  Management  (SWIM)  Section. 

O'Connor,  D.J.  and  Thomann,  R.V.  (1972).  Water  quality  models:  chemical,  physical  and 
biological  constituents.  In:  Estuarine  Modeling:  an  Assessment.  EPA  Water  Pollution 
Control  Research  Series  16070  DZV,  Section  702/71. 

Oliveria,  J.,  Burnett,  W.C.,  Mazzilli,  B.P.,  Brags,  E.S.,  Farias,  L.A.,  Christoff,  J.,  and 
Furtado,  V.  V.  (2003).  Reconnaissance  of  submarine  groundwater  discharge  at  ubatuba  coast, 
Brazil,  using  rn-222  as  a  natural  tracer,  Journal  of  Radioact.,  69:37-52. 


320 

Orth,  R.J.  and  Moore,  K.A.  (1983).  Chesapeake  Bay:  An  unprecedented  decline  in 
submerged  aquatic  vegetation.  Science,  v.  222,  No.  4619:  51-53. 

Phillips,  N.A.  (1957).  A  coordinate  system  having  some  special  advantage  for  numerical 
forecasting.  Journal  of  Meteor.  14:184-185. 

Pickard  ,  G.  L  .  and  Emery,  W.  J.  (1990).  Descriptive  physical  Oceanography.  Pergamon 
Press,  New  York.  Fifth  (SI)  Enlarged  Edition:  320. 

Pribble,  J.R.,  Wade,  D.L.,  Squires,  A.D.,  and  Janicki,  A.J.  (1998).  A  mechanic  water  quality 
model  for  the  tidal  Peace  and  Myakka  rivers.  Proceedings  of  the  Charlotte  Harbor  public 
conference  and  technical  symposium,  Technical  Report  No.  98-02:  242-258. 

Prieur,  L.  and  Sathyendranath,  S.  (1981).  An  optical  classification  of  coastal  and 
oceanographic  waters  based  on  the  specific  spectral  absorption  curves  of  phytoplankton 
pigments,  dissolved  organic  matter,  and  other  particulate  materials.  Limnology  and 
Oceanography,  26(4):67 1-689. 

Rao  P.S.C.,  Jessup,  R.E.,  Reddy,  K.R.  (1984).  Simulation  of  nitrogen  dynamics  in  flooded 
soils.  Soil  Science  138:54-62. 

Reddy,  K.R.,  Jessup,  R.E.,  Rao,  P.S.C.  (1988).  Nitrogen  dynamics  in  a  eutrophic  lake 
sediment.  Hydrobiologia  159:177-188. 

Reddy,  K.R.,  and  Patrick,  W.H.  (1984).  Nitrogen  transformations  and  loss  in  flooded  soil  and 
sediments.  CRC  Critical  Reviews  in  Environmental  Control  13(4):273-309. 

Redfield,  A.,  Ketchman,  B.,  and  Richar,  F.  (1966).  The  influence  of  organisms  on  the 
composition  of  sea-water.  The  Sea  VolII.  Interscience  Publishers,  New  York,  26-48. 

Reed,  R.  K.  (1977).  On  estimating  insolation  over  the  ocean.  Journal  of  Physical 
Oceanography,  7, 482-485.  Research  49(3):  227-239. 

Rutkowski,  C.W.,  Burnett,  W.C.,  Iverson,  R.L.,  and  Chanton,  J.P.  (1999).  The  effect  of 
groundwater  seepage  on  nutrient  delivery  and  seagrasss  distribution  in  the  northern  Gulf  of 
Mexico.  Estuaries,  22:2033-1040. 

Schropp,  S.J.  (1998).  Charlotte  Harbor  Sediment  Quality  Data  Review  and  Evaluation. 
Proceedings  of  the  Charlotte  Harbor  public  conference  and  technical  symposium,  Technical 
Report  No.  98-02:  35-47. 

Sheng,  Y.P.  (1983).  Mathematical  modeling  of  three-dimensional  coastal  currents  and 
sediment  dispersion:  model  development  and  application.  Coastal  Engineering  Research 
Center  Tech.  Rep.  CERC-TR-83-2.  U.S.  Army  Corps  of  Engineers. 


321 

Sheng,  Y.P.  (1986).  Modeling  bottom  boundary  layers  and  cohesive  sediment  dynamics. 
In:(A.J.  Mehta,  ed.)  Estuarine  Cohesive  Sediment  Dynamics.  Pp.360-400.  Springer- Verlag. 
Berlin. 

Sheng,  Y.P.  (1987).  On  modeling  three-dimensional  estuarine  and  marine  hydrodynamics. 
In  Nihoul  and  Jamart,  editors,  Three-dimensional  models  of  marine  and  estuarine  dynamics, 
pages  35-54.  Elswvier,  Amsterdam. 

Sheng,  Y.P.  (1989).  Evolution  of  a  3-D  curvilinear  grid  hydrodynamic  model:  CH3D. 
In:(M.L.  Spaulding,  ed.)  Proceedings  of  the  Estuarine  and  Coastal  Modeling  Conference. 
ASCE,  Newport,  R.I.,  pp.40-49. 

Sheng,  Y.P.  (1994).  Modeling  hydrodynamics  and  water  quality  dynamics  in  shallow  waters. 
In  Proceedings  of  the  first  International  Symposium  on  Ecology  and  Engineering,  Taman 
Negara,  Malaysia.  University  of  Western  Australia  and  Malaysian  Technical  University. 

Sheng,  Y.P.  (1995).  On  Modeling  circulation  in  Florida  Bay.  In  American  Geophysical 
Union  Conference  on  Circulation  in  Intra- American  Seas,  Pueto  Rico.  American  Geophysical 
Union. 

Sheng,  Y.P.  (1998).  Circulation  in  the  Charlotte  Harbor  estuarine  system.  Proceedings  of  the 
Charlotte  Harbor  public  conference  and  technical  symposium,  Technical  Report  No.  98-02: 
91-110. 

Sheng,  Y.P.  (2000).  A  framework  for  integrated  modeling  of  coupled  hydrodynamic- 
sedimentary-ecological  processes.  In  Estuarine  and  Coastal  Modeling,  Volume  6,  pages  350- 
362.  American  Society  of  Civil  Engineers. 

Sheng,  Y.P.,  Ahn,  K.M.,  and  Choi,  J.K.  (1990).  Wind-wave  hindcasting  and  estimation  of 
bottom  sheer  stress  in  Lake  Okeechobee.  Technical  Report  UFL/COEL-93/02 1 ,  Coastal  and 
Oceanographic  Engineering  Department,  University  of  Florida,  Gainesville,  Florida. 

Sheng,  Y.P.,  Chen,  X.J.,  Schofield,  S.,  and  Yassuda,  E.  (1993).  Hydrodynamics,  sediment 
and  phosphorous  dynamics  in  Lake  Okeechobee  during  an  episodic  event.  Technical  Report, 
Coastal  and  Oceanographic  Engineering  Department,  University  of  Florida,  Gainesville, 
Florida. 

Sheng,  Y.P.,  and  Chiu,  S.S.  (1986).  Tropical  cyclone  generated  currents.  In  Proceedings  of 
the  20th  International  Conference  on  Coastal  Engineering,  pages  737-751.  American  Society 
of  Civil  Engineers. 

Sheng,  Y.P.,  Christian,  D.,  and  Kim,  T.Y.  (2002c).  A  IRL  light  attenuation  model.  Indian 
River  Lagoon  Pollutant  Load  Reduction  (IRLPLR)  Model  Development  Project.  Final  Report 
Volume  II  (A  3-D  Water  Quality  Model)  to  St.  Johns  River  Water  Management  District. 
Coastal  and  Oceanographic  Engineering  Department,  University  of  Florida,  Gainesville, 
Florida. 


322 

Sheng,  Y.P.,  Davis,  J.R.,  and  Park,  K.  (2003)  Development  of  parallel  computing  version  of 
CH3D:  Parallel  CH3D  User's  Manual.  Amendment  to  the  three  dimensional  water  quality 
model  of  the  Charlotte  Harbor  estuarine  system  for  the  South  Florida  Water  Management 
District. 

Sheng,  Y.P.,  Davis,  J.R.,  Sun,  D.,Qiu,  C,  Park,  K.,  Kim,  T.Y.,  and  Zhang,  Y.  (2001). 
Application  of  an  integrated  modeling  system  of  estuarine  and  coastal  ecosystem  to  Indian 
River  Lagoon,  Florida.  In  Estuarine  and  Coastal  Modeling,  Session  4B.  American  Society 
of  Civil  Engineers,  pp. 329-343 

Sheng,  Y.P.,  and  Lick,  W.  (1979).  Resuspension  and  transport  of  sediments  in  a  shallow 
lake.  Journal  of  Geophysical  Research,  V.  84,  pp.  1809-1826. 

Sheng,  Y.P.  and  Park,  K.  (2000).  A  3-D  Circulation  and  Salinity  Model  of  Charlotte  Harbor 
Estuarine  System.  Phase  I  Final  Report  to  Southwest  Florida  Water  Management  Department 
(SFWMD),  University  of  Florida,  Gainesville,  Florida. 

Sheng,  Y.P.  and  Park,  K.  (2001).  Modeling  the  Effects  of  Caloosahatchee  Discharges  on 
Circulation  and  Salinity  in  Charlotte  Harbor.  Interim  Report  Southwest  Florida  Water 
Management  Department  (SFWMD),  University  of  Florida,  Gainesville,  Florida. 

Sheng,  Y.P.  and  Park,  K.  (2002).  3-dimensional  Water  Quality  model  for  Charlotte  Harbor 
estuarine  system.  Phase  HI  Final  Report  Southwest  Florida  Water  Management  Department 
(SFWMD),  University  of  Florida,  Gainesville,  Florida. 

Sheng,  Y.P.,  Qiu,  Q.,  Park,  K.,  and  Kim,  T.Y.  (2002b).  A  3-D  water  quality  model.  Indian 
River  Lagoon  Pollutant  Load  Reduction  (IRLPLR)  Model  Development  Project.  Final  Report 
Volume  II  (A  3-D  Water  Quality  Model)  to  St.  Johns  River  Water  Management  District. 
Coastal  and  Oceanographic  Engineering  Department,  University  of  Florida,  Gainesville, 
Florida. 

Sheng,  Y.P.,  Sun,  D.,  and  Zhang,  Y.  (2002a).  A  3-D  IRL  suspended  sediment  transport 
model.  Indian  River  Lagoon  Pollutant  Load  Reduction  (IRLPLR)  Model  Development 
Project.  Final  Report  to  St.  Johns  River  Water  Management  District.  Coastal  and 
Oceanographic  Engineering  Department,  University  of  Florida,  Gainesville,  Florida. 

Sheng,  Y.P.,  and  Villaret,C.  (1989).  Modeling  the  effect  of  suspended  sediment  stratification 
on  bottom  exchange  processes.  Journal  of  Geophysical  Research  -  Oceans,  94(C10):  14229- 
14444. 

Simon,  N.S.  (1989).  Nitrogen  cycling  between  sediment  and  the  shallow-water  column  in 
the  transition  zone  of  the  Potomac  River  and  Estuary.  U.  The  role  of  wind-driven 
resuspension  an  adsorbed  ammonium.  Estuarine,  Coastal  and  Shelf  Science  28:531-547. 

Smith,  R.C.  and  Baker,  K.  (1981).  Optical  properties  of  the  clearest  waters.  Applied  Optics, 
20(2):  177-184. 


323 

Smith,  S.  D.  (1980).  Wind  stress  and  heat  flux  over  the  ocean  in  gale  force  winds.  Journal 
of  Physical  Oceanography,  10,  709-726 

Smith,  S.  D.  (1989).  Coefficients  for  sea  surface  wind  stress,  heat  flux,  and  wind  profiles  as 
a  function  of  wind  speed  and  temperature.  Journal  of  Geophysics  Research,  93, 15467-15472 

Steele,  J. A.  (1974).  The  structure  of  marine  ecosystem.  Oxford,  Blackwell:  p74-91. 

Steele,  J.M.  (1965)  Notes  on  some  theoretical  problems  in  production  ecology.  In:(C.R. 
Goldman  ed.)  Primary  Production  in  Aquatic  Environments.  University  of  California  Press. 
Berkeley 

Steinberger,  N.,  and  Hondzo,  M.  (1999).  Experimental  study  of  the  hydrodynamic  control 
of  diffusional  mass  transfer  at  the  sediment-water  interface.  Journal  of  Environmental 
Engineering,  125(2):  192-200 

Stoker,  Y.E.  (1986).  Water  quality  of  the  Charlotte  Harbor  estuarine  system,  Florida, 
November  1982  through  October  1984.  U.S.  Geological  Survey  Open-File  Report  85-563: 

213. 

Stoker,  Y.E.  (1992).  Salinity  distribution  and  variation  with  fresh-water  inflow  and  tide,  and 
potential  changes  in  salinity  due  to  altered  freshwater  inflow  in  the  Charlotte  Harbor 
estuarine  system,  Florida.  U.S.  Geological  Survey  Water-Resources  Investigations  Report 
92-4062:  30. 

Stull,  R.  B.  (1988).  An  Introduction  to  Boundary  Layer  Meteorology.  Kluwer  Academic 
Publishers,  Netherlands,  660pp. 

Sun,  D.  and  Sheng,  Y.P.  (2001).  Modeling  suspended  sediment  transport  under  combined 
wave  current  interactions  in  Indian  River  Lagoon.  PhD  dissertation,  University  of  Florida, 
Gainesville,  Florida. 

Taylor,  P.,  Gulev,  S.,  Barnier,  B.,Bradiey,  F.,Charlock,  T.,Gleckler,  P.,Kubota,  M., 
Kutsuwada,  M.,Legler,  D.,  Lindau,  R.,  Rothrock,  D.,  Schulz,  J.da-Silva,  A.,  Sterl,  A.,  and 
White,  G.  (2000).  Intercomparison  and  validation  of  ocean-atmosphere  energy  flux  fields. 
Final  Report  of  the  joint  WCRP/SCOR  working  group  on  air-sea  fluxes,  WCRP- 112,  WMO/ 
TD-No.  1036,  303  S 

Thomann,  R.V.  (1992).  Section  14.1.  in  Technical  guidance  manual  for  performing  waste 
load  allocations.  Book  III:  Estuaries-Part  4,  Critical  review  of  coastal  embayment  and 
estuarine  waste  load  allocation  modeling.  U.S.  EPA  Technical  Report  EPA-823-R-92-005. 
Washington,  D.C. 

Thomann,  R.  V.,  and  Fitzpatrick,  J.F.  (1982).  Calculation  and  verification  of  a  amthematical 
model  of  the  eutrophication  of  the  Potomac  Estuary.  Prepared  for  the  Department  of 
Environmental  Services,  Government  of  the  Distric  of  Columbia. 


324 

Thompson,  J.F.  (1982).  Numerical  Grid  Generation,  chapter  General  Curvilinear  Coordinate 
Systems,  pages  1-30.  Elsevier,  Amsterdam. 

Thompson,  J.F.,  Warsi,  Z.U.A.  and  Mastin,  C.W.  (1985).  Numerical  Grid  Generation: 
Foundations  and  Applications.  North-Holland.  New  York 

Weast  (1977).  Handbook  of  chemistry  and  physics,  58th  ed.,  Table  F-200 

Witkowski,  J.  and  Jaffe,  P.R.  (1987).  Kinetic  and  equlibrium  formulations  in  modeling  of 
hydrophobic  organic  chemicals  in  surface  and  ground  waters.  System  Analysis  in  Water 
Quality  Management  (M.B.  Beck,  ed.),  Advences  in  Water  Pollution  Control-  Aseries  of 
Conferences  Sponsored  by  IAWPRC. 

Wolfe,  S.H.,  and  Drew  R.  D.,  eds.  (1990).  An  ecological  characterization  of  the  Tampa  Bay 
watershed.  U.S.  Fish  Wildl.  Serv.  Biol.  Rep.  90:  20. 

Yassuda,E.A.  andSheng,  Y.P.  (1996).  Integrated  modeling  of  Tampa  Bay  Estuarine  System. 
PhD  dissertation,  University  of  Florida,  Gainesville,  Florida. 

Zison,  S.,  Mills,  W.B.,  Deimer,  D.,  and  Chen,  C.W.  (1978).  Rates,  constants  and  kinetic 
formulations  in  surface  water  quality  modeling.  Prepared  by  Tetra  Tech,  Inc.,  Lafayette,  CA 
for  Environmental  Research  Laboratory,  U.S.  Environmental  Protection  Agency,  Athen,  GA. 
EPA-600/3-78-105:335. 


BIOGRAPHICAL  SKETCH 
Kijin  Park  was  born  in  Pusan,  the  largest  harbor  of  Korea,  on  the  25th  of  October.  He 
received  the  Bachelor  of  Science  degree  and  M.S.  degree  in  department  of  Marine  Science 
at  the  Pusan  National  University  in  February,  1991  and  1994,  respectively.  After  finished 
M.S.  degree,  he  had  worked  as  research  scientist  at  Korea  Oceanography  Research  & 
Development  Institute  (KORDI)  and  Korea  Meteorological  Administration  (KMA)  until 
June,  1997.  The  following  summer  he  left  for  Gainesville,  Florida  ,  and  began  his  graduate 
studies  in  coastal  and  oceanographic  engineering  at  the  University  of  Florida. 


325 


I  certify  that  I  have  read  this  study  and  that  in  my  opinion  it  conforms  to 
acceptable  standards  of  scholarly  presentation  and  is  fully  adequate,  in  scope  and  quality, 
as  a  dissertation  for  the  degree  of  Doctor  of  Philosophy. 


Y.  Pe^fSneng,  Chair^ 

Professor  of  Civil  and  Coastal  Engineering 

I  certify  that  I  have  read  this  study  and  that  in  my  opinion  it  conforms  to 
acceptable  standards  of  scholarly  presentation  and  is  fully  adequate,  in  scope  and  quality, 
as  a  dissertation  for  the  degree  of  Doctor  of  Philosophy. 

Robert  G.  Dean 

Graduate  Research  Professor  of  Civil  and 
Coastal  Engineering 

I  certify  that  I  have  read  this  study  and  that  in  my  opinion  it  conforms  to 
acceptable  standards  of  scholarly  presentation  and  is  fully  adequate,  in  scope  and  quality, 
as  a  dissertation  for  the  degree  of  Doctor  of  Philosophy. 


Robert  J.  Thiek£ 

Assistance  Professor  of  Civil  and  Coastal 
Engineering 

I  certify  that  I  have  read  this  study  and  that  in  my  opinion  it  conforms  to 
acceptable  standards  of  scholarly  presentation  and  is  fully  adequate,  in  scope  and  quality, 
as  a  dissertation  for  the  degree  of  Doctor  of  PhilcrStoph\ 


Louis  H.  Motz 

Associate  Professor  of  Civil  and  Coastal 
Engineering 


I  certify  that  I  have  read  this  study  and  that  in  my  opinion  it  conforms  to 
acceptable  standards  of  scholarly  presentation  and  is  fully  adequate,  in  scope  and  quality, 
as  a  dissertation  for  the  degree  of  Doctor  of  Philosopl; 


K.  Ramesh  Reddy 

Graduate  Research  Professor  of  Soil  and 
Water  Science 


This  dissertation  was  submitted  to  the  Graduate  Faculty  of  the  College  of 
Engineering  and  to  the  Graduate  School  and  was  accepted  as  partial  fulfillment  of  the 
requirements  for  the  degree  of  Doctor  of  Philosophy. 


)  Ajgw^l  t/j^-t^-yv-cW^ 


August,  2004 

Pramod  P.  Khargonekar 
Dean,  College  of  Engineering 


Kenneth  J.  Gerhardt 

Interim  Dean,  Graduate  School 


,PA3&> 


UNIVERSITY  OF  FLORIDA 


3  1262  08554  5670