Figure 2 - uploaded by Robert Weisberg
Content may be subject to copyright.
The model grid showing the along-axis (I) and across-axis (II) sections (bold lines) chosen for analyses and the tide gauge (solid circles) and current meter (solid triangle) stations for model simulation comparisons. The square denotes a site located just inside of Boca Grande Pass at which we present tidal current profiles. 

The model grid showing the along-axis (I) and across-axis (II) sections (bold lines) chosen for analyses and the tide gauge (solid circles) and current meter (solid triangle) stations for model simulation comparisons. The square denotes a site located just inside of Boca Grande Pass at which we present tidal current profiles. 

Source publication
Article
Full-text available
The circulation of the Charlotte Harbor (CH) estuary is explored with a primitive equation model that encompasses the estuary and the adjacent West Florida Shelf. Tidal forcing is from the shelf through the inlets. We use the M2, S2, K1, and O1 constituents that account for 95% of the shelf tidal variance. River inflows are by the Peace, Myakka, an...

Contexts in source publication

Context 1
... The initial decrease across the inlet is by dissipation. The subsequent increase is by wave reflection at the estuary’s head. Because of the narrowness of the inlet and dissipation a sufficient pressure head (by Bernoulli’s theorem) is required to drive water through the inlet. This is achieved by the phase gradient across the inlet. Thus the M 2 , S 2 , K 1 , and O 1 constituents’ phases increase across the inlet by about 35 ° , 45 ° , 20 ° , and 20 ° , respectively. They continue increasing toward the Peace River by amounts larger p ffiffiffiffiffi than implied for a gravity wave propagating at speed gh , which is consistent with frictional loses within the estuary [ Friedrichs and Madsen , 1992]. The amplitude and phase behaviors into the estuary from the SCB entrance are somewhat different. The amplitudes decrease monotonically because of the shoaling depths into Matlacha Pass and Pine Island Sound (a frictional effect). They also decrease into the Caloosahatchee River toward Ft. Myers because of both friction and the several right angle bends in the channel, each requiring a drop in pressure. For example, the amplitude of the M 2 tide drops some 12 cm between the SCB entrance and Ft. Myers, and the phase advances substantially. [ 16 ] The model-predicted surface currents at the maximum flood and ebb phases of a spring tide are shown in Figure 5. At maximum flood, the currents over the innershelf are weak. As water converges on the inlets the currents accelerate and are much stronger. They are a maximum at the BGP, attaining speeds of about 1.4 m s À 1 . Currents through the SCB entrance are smaller since that inlet is wider. These currents flow into Matlacha Pass, Pine Island Sound, and the Caloosahatchee River. At maximum ebb, the currents are reversed and stronger (except in the regions of Pine Island Sound and Matlacha Pass). At BGP, for example, the maximum ebb current is 1.5 m s À 1 . Consistent with the observations of Goodwin [1996], these findings suggest that the CH estuary is slightly ebb dominant. Further evidence of ebb dominance (not shown) derives from the durations of the ebb and flood tides. Sampled along sections from BGP to the Peace River the flood phase is found to exceed the ebb phase of the tide by approximately 2.0 to 0.5 hours. The ebb-flood asymmetry is much less during neap tides (Figure 6), and this finding is consistent with other estuaries as reported by Aubrey and Speer [1985]. With asymmetry deriving from nonlinear interaction it follows that asymmetry should decrease with decreasing amplitude from spring to neap tide. [ 17 ] Strong tidal currents induce vertical velocity variations at the passes. Figure 6 shows spring and neap tide vertical velocity distributions at BGP for the slack high, slack low, and the maximum ebb and flood phases. Max- imum ebb (flood) shows downwelling (upwelling) at the location sampled just inside the pass with an ebb-flood asymmetry and with the spring tide values about twice those at neap tide. These behaviors may be explained by the topographic variations along the channel axis (with maximum depth within the inlet) and the bottom kinematic boundary condition. [ 18 ] Following the procedure of section 3 river discharges are switched on [as a volume flux boundary condition [ Chen et al. , 1999] after 10 M 2 cycles and then held constant for 120 M 2 cycles to allow the model-predicted salinities to reach a quasi-steady state. We use the spring 1998 mean discharges of 165, 40, and 240 m 3 s À 1 for the Peace, Myakka, and Caloosahatchee Rivers, respectively. Since spring 1998 was anomalously wet these values are meant to reflect conditions of large river inflows. In this section we show results that are averaged over the next two M 2 cycles to explore the nontidal circulation and salinity fields determined by the combined effects of tides and rivers. [ 19 ] The nontidal currents are strongly affected by rivers since (with tidal mixing) these impart the gravitational convection portion of the circulation that is generally referred to as estuarine circulation [e.g., Cameron and Pritchard , 1963]. Figure 7 shows the tide-averaged surface current and salinity fields. Waters from the Peace and Myakka Rivers converge, and the combined flow then proceeds southward down the main body of the estuary. Note that the tide-averaged currents are stronger on the western side, as is also reflected by the lower-salinity values there. A model twin experiment (not shown) omitting the Coriolis acceleration gave different results, with the current and salinity fields being more symmetric about the estuary axis and driven by a concave sea surface height due to the centrifugal acceleration. This twin experiment demonstrates that the right hand side preference when looking down- stream is a consequence of the Coriolis acceleration. [ 20 ] From both the velocity and salinity patterns we see that the estuary partitions into two units, a CH side, influenced by the Peace and Myakka Rivers, and a SCB side, influenced by the Caloosahatchee River. Separating these two units are the shallow Matlacha Pass and Pine Island Sound regions. Effectively, the system consists of two nearly independent estuaries. Fresh water from Caloosahatchee River exits primarily through the mouth of SCB, whereas fresh water from the Peace and Myakka Rivers exits primarily through BGP. [ 21 ] Once on the innershelf waters emanating from the CH estuary are affected by the local topography and the Coriolis acceleration with the outflows sweeping in broad anticyclonic arcs. The salinity field reflects this behavior, with a near surface lens of relatively low-salinity water organizing into a northward-directed current. [ 22 ] Returning to the estuary, a sampling of the tide- averaged circulation and salinity in a vertical plane oriented along the main channel axis (the bold line I in Figure 2) reveals a two-layered net estuarine circulation (Figure 8) with fresher water flowing seaward atop saltier water flowing landward, as expected by gravitational convection. The maximum down-estuary current (within the upper 1 m of the water column) is about 0.12 m s À 1 , compared with the maximum up-estuary current of 0.06 m s À 1 . On cross- sectional average, the net transport equals the river discharge rate. Note that before diminishing and reversing sign along this section the outflowing water turns right to exist through BGP. [ 23 ] The vertical structures of the estuarine circulation and the salinity stratification are largely determined by tidal mixing. Since tidal mixing originates from friction along the bottom the salinity isolines of Figure 8 are nearly vertical there, as contrasted with the isolines being more nearly horizontal at the surface. [ 24 ] Estuaries are also influenced by winds, either directly by the stress on the estuary surface [ Weisberg and Sturges , 1976; Weisberg , 1976], or indirectly by wind stress effects on the adjacent coastal ocean [ Wang , 1979]. We begin to address this for the CH estuary by forcing the model with spatially uniform winds directed either toward the northwest or the southeast. Typical of winds for the region we use a speed of 5 m s À 1 . Adding to section 4 the model is spun up to tides over 10 M 2 cycles, then to rivers over 120 M 2 cycles, and then to the uniform winds over another 10 M 2 cycles. [ 25 ] The wind directions are intended to explore three factors. First, with respect to the WFS, southeasterly (northwesterly) winds are downwelling (upwelling) favorable. They cause onshore (offshore) Ekman transports and along-shore currents directed toward the northwest (southeast) [e.g., Li and Weisberg , 1999]. Second, with respect to the estuary, they act in opposition to (or in concert with) the two-layered estuarine circulation induced by gravitational convection (Figure 8). Third, by adding an additional source for mixing from the surface down to go along with the tidal mixing from the bottom up the winds largely alter the estuary’s salinity ...
Context 2
... 10. Along-axis (section I of Figure 2) tide- averaged (top) current and (bottom) salinity distributions for the case of tides, plus rivers, plus a 5 m s À 1 downwelling favorable wind. The contour intervals on currents and salinity are 3 cm s À 1 and 1 psu, respectively.  ...
Context 3
... Figure 15. Up-estuary salt flux integrated across section II of Figure 2: (top) Tides and rivers only, and (bottom) tides, plus rivers, plus 5 m s À 1 oscillatory winds. Three lines are shown for either case. The solid, time-varying lines are the low-pass-filtered (nontidal) salt flux that consists of contributions by the mean gravitational convection (the dashed lines) and the Reynolds’ fluxes due to rectification by either the tides or by the tides plus winds (the difference between the dashed line and the dot-dashed ...
Context 4
... section as saltier water upwells and proceeds back toward the south tending to override fresher water. [ 30 ] The increase in the two-layered flow, and conse- quently the increase in the horizontal advection of salinity, maintains the vertical stratification even in the presence of increased mixing by winds, although this stratification is weaker than that in the case without winds ( Figure 8). With a vertical salinity gradient extending down to the bottom and stabilizing the water column there, the tidal mixing is also suppressed. [ 31 ] In either case (southeasterly, downwelling favorable or northwesterly, upwelling favorable) we see that spatially uniform, steady winds largely alter the circulation and salinity structures of the CH estuary. Given these findings we next look at the effects of oscillatory winds typical of extratropical weather propagating through the region. [ 32 ] As in section 5 we use a wind speed of 5 m s , but instead of maintaining constant southeasterly or northwesterly winds we allow the winds to oscillate about a zero mean value with periodicities of 4, 8, and 16 M 2 cycles (Figure 13). These oscillatory winds are added after the model is spun up for tides over 10 M 2 cycles and for tides, plus rivers over another 120 M 2 cycles. The model is then run for 10 wind cycles (i.e., 40, 80, and 160 M 2 cycles for wind periods of 4, 8, and 16 M 2 cycles, respectively), after which the outputs are filtered to remove oscillations occurring on timescales shorter than 36 hours and then averaged over the full 10 wind cycles. We give results at the times of the maximum upwelling and downwelling wind phases and averaged over all cycles. [ 33 ] At the maximum downwelling favorable phase of the wind cycle (Figure 14a), the near-surface speed decreases with increasing wind period. For example, the maximum upstream flow speed of about 0.24 m s À 1 for the 4 M 2 cycle wind period, decreases to 0.12 m s À 1 and 0.09 m s À 1 for the 8 and 16 M 2 cycle wind periods, respectively. The salinity structure is also related to wind period, with the vertical stratification decreasing with increasing wind period. These results are consistent with the findings under constant winds. Downwelling favorable winds oppose the gravitational convection. For short-period oscillations the upstream current can accelerate with little opposition by an ensuing surface pressure gradient force. With increasing period the surface pressure gradient can set up to oppose the acceler- ating surface current, thereby reducing its magnitude. Similarly, increasing the wind period provides more time to transport higher-salinity water from the lower to the higher reaches of the estuary, which causes the salinity isolines to rotate more vertical. [ 34 ] The vertical structure of the circulation and salinity for the maximum upwelling favorable winds are also a function of wind period (Figure 14b). The along-axis flow magnitude decreases with increasing period in part due to an opposing surface pressure gradient force, but also due to a decrease in the baroclinic portion of the pressure gradient force as the salinity isolines become more horizontal. The direct affects of the wind on both the barotropic (surface slope) and the baroclinic (isohaline slope) portions of the pressure gradient largely outweigh those by gravitational convection alone (the no wind case). In the CH, as in other estuaries, winds, even of modest magnitude, can dominate the tide-averaged flow and the resulting material property fields. [ 35 ] Given that the tide-averaged, along-axis velocity and salinity fields are sensitive to the oscillatory wind period it is natural to ask whether or not the tide and oscillatory wind averaged net estuarine circulation and salinity fields are. As shown by Figure 14c, this sensitivity to the oscillatory wind period is much less perceptible. All three along-axis velocity and salinity panels look alike with a two-layered estuarine circulation and salinity stratification somewhat similar to the no wind case of Figure 8. Because of the added mixing by the winds over that by the tides and rivers alone the vertical salinity gradient is less, as is the magnitude of the gravitational convection. This begs the question of what determines the magnitude of the estuarine circulation under the combined effects of tides, rivers, and winds, each of which contribute to the averaged distributions of pressure gradient driving and frictional retarding forces. We treat this topic in a separate study on the energetics of the CH estuary [ Weisberg and Zheng , 2003], which shows that the net estuarine circulation at first increases with increasing mixing and then decreases when the rate of energy input to turbulence production exceeds the rate of work against buoyancy. [ 36 ] Oscillatory winds and tides also affect net material transports. We illustrate this by considering the up-estuary salt flux integrated across section II of Figure 2. Compara- tive results are given for the cases of tides and rivers and tides, rivers, and 5 m s À 1 oscillatory winds (Figure 15 (top) and 15 (bottom), respectively). We first note that the instantaneous salt flux varies on all forcing function timescales. Here we calculate a nontidal up-estuary salt flux (using a 36 hour low-pass filter on the integrated product of the up- estuary velocity component and salinity), which we then further average over 80 M 2 cycles to get the mean and the horizontal Reynolds’ contributions. We define the mean contribution as the integrated product of the mean up-estuary velocity component and the mean salinity by gravitational convection, and we define the horizontal Reynolds’ contribution as the integrated product of the deviations about these means. The record length mean of the nontidal salt flux is the sum of these two contributions. [ 37 ] Without winds we find a spring to neap tide modulation in the nontidal salt flux, and upon averaging this we see that the record length mean salt flux is more than twice as large as the salt flux by the mean gravitational convection. The difference between these two is the Reynolds’ flux, in this case due to tidal rectification [e.g., Lewis and Lewis , 1983; Geyer and Nepf , 1996]. With winds we find a nontidal salt flux modulation on the timescales of both the winds (in this case with an 8 M 2 cycle periodicity) and the spring to neap tide cycle, but only a slightly larger difference (the Reynolds’ contribution) between the record length mean salt flux and the portion by gravitational convection. From these comparative findings we conclude that the Reynolds’ salt flux is primarily a consequence of tidal rectification. While not shown, increasing the wind speed increases the Reynolds’ contribution, but decreases the contribution by gravitational convection by a slightly larger amount such that the record length mean salt flux decreases. This arises because the mean circulation by gravitational convection [ Weisberg and Zheng , 2003] and the correlation between the fluctuating velocity and salinity values are both sensitive to wind speed. [ 38 ] The circulation of the CH estuary, driven by tides, rivers, and winds, is explored using a three-dimensional, primitive equation model. Tidal forcing is from the shelf through the inlets, and we consider the M 2 , S 2 , K 1 , and O 1 constituents that account for some 95% of the WFS tidal variance. River discharges include the spring 1998 mean values for the Peace, Myakka, and Caloosahatchee ...
Context 5
... Clima- tology [e.g., Hellerman and Rosenstein , 1983] shows pre- vailing offshore winds (northeasterly to easterly) from September to February, followed by a spring transitional regime (March to May) in which the winds shift from easterly to southeasterly. Weak southeasterly winds then prevail from June to August. The climatology is punctuated by synoptic-scale variability as extratropical weather systems propagate across the region in fall through spring. During spring and summer the winds are also impacted by sea breeze and tropical storms. Generally, with the excep- tion of tropical storms the winds tend to be strongest in response to the extratropical weather patterns from fall through spring. [ 6 ] Material properties respond to the variations in rivers, tides, and winds. Seasonal salinity variations occur primarily in response to seasonal river discharge variations, whereas daily salinity variations occur primarily in response to tides, with highest (lowest) salinity values occurring at slack high (low) water. Observations (E. Estevez, Mote Marine Laboratory, personal communication) show that the CH estuary may be classified as partially to well-mixed except during high river discharge periods when the water column may be highly stratified. [ 7 ] A model study by Goodwin [1996] demonstrated agreements for tidal currents and sea level when gauged against data. That model, being two-dimensional (vertically averaged) and exclusive of winds, was neither capable of describing the salinity structure, nor the effects of wind on the stratification or flow fields. Thus the physical mecha- nisms that drive the nontidal, density-driven estuarine circulation and the wind-driven variations of the circulation remain to be described for the CH estuary. Since these aspects of the circulation largely control both the stratification and the related state variable distributions, a more complete understanding of the circulation is necessary to support multidisciplinary ecological studies. [ 8 ] Here we use a three-dimensional, primitive equation model to investigate the tide, buoyancy, and wind-driven circulations of the CH estuary and the exchanges of water that occur between the estuary and the adjacent WFS. Section 2 describes the model and the experimental design. Section 3 presents the tidal portion of the model simulations. River discharges are added in section 4 to examine the buoyancy effects on the circulation and salinity fields. Constant winds in either downwelling or upwelling favorable directions are then added in section 5 to examine the roles of all three forcing elements on the circulation and salinity fields. Since winds tend to be oscillatory in nature, we examine this effect in section 6, and the results are summarized in section 7. [ 9 ] We use the primitive equation, ECOM3D-si model, based on the Princeton Ocean Model (POM) of Blumberg and Mellor [1987] as subsequently modified by Blumberg [1993], Chen and Beardsley [1995], and Zheng et al. [2003]. The model incorporates the Mellor and Yamada [1982] level 2.5 turbulence closure scheme to provide flow- dependent turbulent mixing parameters along with a free surface. A s coordinate in the vertical and an orthogonal curvilinear coordinate in the horizontal are used. Unlike POM that employs a mode-splitting technique, ECOM3D-si uses a semi-implicit finite difference scheme in which the advection, Coriolis, baroclinic pressure gradient, and horizontal diffusion terms are calculated explicitly, and the barotropic pressure gradient (or surface elevation gradient) and the vertical diffusion terms are solved for implicitly. Advantages of this semi-implicit scheme are: (1) it gener- ates a symmetric, positive definite set of equations for sea level that can be solved by a preconditioned conjugate gradient method [ Casulli , 1990] and (2) it removes the external CFL constraint allowing for larger time steps by eliminating the need for separate internal and external modes. Successful applications of ECOM3D-si exist for ocean [e.g., Mellor and Ezer , 1991], coastal [e.g., Blumberg et al. , 1993; Chen and Beardsley , 1995; Chen et al. , 1999], and estuarine [e.g., Blumberg and Pritchard , 1997; Zheng et al. , 2003] environments. [ 10 ] A common problem to the salinity equation, when central differences are used for the advection terms, is the occurrence of negative salinity if the estuary is shallow and the river discharge is large. To avoid this we employ the Multidimensional Positive Definite Advection Transport Algorithm [ Smolarkiewicz , 1984] that uses an ‘‘antidiffu- sion’’ velocity in a successive upwind scheme to correct first-order truncation errors, yielding a positive definite, second-order accurate advection scheme. Successful applications of this scheme exist for nutrient transport studies on Georges Bank [ Chen and Beardsley , 1998] and for salinity simulations in the Satilla River estuary [ Zheng et al. , 2003]. [ 11 ] The model grid is shown in Figure 2, with an open boundary located on the shelf some 30 km away from the barrier islands. A shelf component is necessary for two reasons. First, with multiple inlets, the only way to specify the estuary’s tidal forcing is to apply a shelf tide model. Second, the shelf component allows us to address the material exchanges between the shelf and the estuary. The model grid extends out to the 25 m isobath, arching to the coast near Venice, Florida in the north and Naples, Florida. in the south. There are 250 (along-shelf) Â 148 (across-shelf) grid points varying in resolution from about 250 m inside the estuary to about 1100 m at the open boundary. Eleven evenly distributed s levels provide vertical resolution of less than 0.1 m in shallow regions to between 1 $ 2 m on the innershelf. We use National Ocean Service bathymetry data with 30 m resolution over most of the domain. In regions such as the Caloosahatchee River and Matlacha Pass, where bathymetry data are not given in this database, we digitized bathymetry directly from NOAA charts. A time step of 124.2 s (equating to 360 time steps per M 2 tide cycle) is used. [ 12 ] Model runs begin from a state of rest with initial salinity and temperature values set at 35 psu and 20 ° C, respectively. At the open boundary sea level is controlled using the M 2 , S 2 , K 1 , and O 1 constituents sampled from the He and Weisberg [2002] WFS tide model, and a radiation p ffiffiffiffiffi boundary condition with a propagation speed of gh is used for velocity. Temperature is held constant throughout, and a nongradient open boundary condition is applied to salinity such that outgoing water may leave with its prognostic salinity, whereas incoming water enters with a salinity of 35 psu. To avoid instability the tidal amplitudes are ramped up from zero to their full values over the first M 2 tide cycle. [ 13 ] Three different classes of experiments are conducted. The first is a barotropic tide simulation. The second includes tides and rivers with prognostic salinity. The third adds either constant or oscillatory winds to the tide, plus river case. [ 14 ] After the initial ramp up over one M 2 tide cycle the model is run to quasi-equilibrium over an additional 9 M 2 cycles. Sixty more M 2 cycles are then run for analysis. We employ a linear least squares harmonic analysis (following Foreman [1977, 1978]) to compute the amplitudes, phases, and current ellipse parameters for the M 2 , S 2 , K 1 , and O 1 constituents at each of the model grid points. The model- predicted tidal elevations agree well with observations at the Naples and Ft. Myers tide gauges (Figure 3). The model duplicates the spring to neap tide variation, including the more diurnal (semidiurnal) character at spring (neap) tide. The simulation at Naples is better than at Ft. Myers since Naples is closer to the open boundary where the tides are forced and Ft. Myers is affected by depth and geometry complexities. We found that by improving the bathymetric data from the Caloosahatchee River we improved the Ft. Myers simulation. Along with these sea level comparisons, Table 1 shows the observed and computed tidal current ellipse parameters at a mooring positioned southwest of Sanibel (see Figure 2). Combined, the sea level and velocity comparisons demonstrate the legitimacy of the open boundary condition in forcing the tides. [ 15 ] Model-predicted coamplitude and cophase charts for the M 2 , S 2 , K 1 , and O 1 tidal constituents are presented in Figures 4a and 4b, respectively. The local innershelf results are consistent with the shelf-wide findings of He and Weisberg [2002]. The semidiurnal species show amplitude gradients from north to south, whereas the diurnal species are spatially more uniform. The tides propagate across the innershelf and into the estuary through BGP, Captiva Pass, and the mouth of SCB. Through BGP, the M 2 , S 2 , K 1 , and O 1 constituent amplitudes first decrease by 2.5 cm, 2 cm, 1 cm, and 1 cm, respectively, before increasing farther north into the Peace River by 6 cm, 3 cm, 1.5 cm, and 2 ...
Context 6
... coordinate in the horizontal are used. Unlike POM that employs a mode-splitting technique, ECOM3D-si uses a semi-implicit finite difference scheme in which the advection, Coriolis, baroclinic pressure gradient, and horizontal diffusion terms are calculated explicitly, and the barotropic pressure gradient (or surface elevation gradient) and the vertical diffusion terms are solved for implicitly. Advantages of this semi-implicit scheme are: (1) it gener- ates a symmetric, positive definite set of equations for sea level that can be solved by a preconditioned conjugate gradient method [ Casulli , 1990] and (2) it removes the external CFL constraint allowing for larger time steps by eliminating the need for separate internal and external modes. Successful applications of ECOM3D-si exist for ocean [e.g., Mellor and Ezer , 1991], coastal [e.g., Blumberg et al. , 1993; Chen and Beardsley , 1995; Chen et al. , 1999], and estuarine [e.g., Blumberg and Pritchard , 1997; Zheng et al. , 2003] environments. [ 10 ] A common problem to the salinity equation, when central differences are used for the advection terms, is the occurrence of negative salinity if the estuary is shallow and the river discharge is large. To avoid this we employ the Multidimensional Positive Definite Advection Transport Algorithm [ Smolarkiewicz , 1984] that uses an ‘‘antidiffu- sion’’ velocity in a successive upwind scheme to correct first-order truncation errors, yielding a positive definite, second-order accurate advection scheme. Successful applications of this scheme exist for nutrient transport studies on Georges Bank [ Chen and Beardsley , 1998] and for salinity simulations in the Satilla River estuary [ Zheng et al. , 2003]. [ 11 ] The model grid is shown in Figure 2, with an open boundary located on the shelf some 30 km away from the barrier islands. A shelf component is necessary for two reasons. First, with multiple inlets, the only way to specify the estuary’s tidal forcing is to apply a shelf tide model. Second, the shelf component allows us to address the material exchanges between the shelf and the estuary. The model grid extends out to the 25 m isobath, arching to the coast near Venice, Florida in the north and Naples, Florida. in the south. There are 250 (along-shelf) Â 148 (across-shelf) grid points varying in resolution from about 250 m inside the estuary to about 1100 m at the open boundary. Eleven evenly distributed s levels provide vertical resolution of less than 0.1 m in shallow regions to between 1 $ 2 m on the innershelf. We use National Ocean Service bathymetry data with 30 m resolution over most of the domain. In regions such as the Caloosahatchee River and Matlacha Pass, where bathymetry data are not given in this database, we digitized bathymetry directly from NOAA charts. A time step of 124.2 s (equating to 360 time steps per M 2 tide cycle) is used. [ 12 ] Model runs begin from a state of rest with initial salinity and temperature values set at 35 psu and 20 ° C, respectively. At the open boundary sea level is controlled using the M 2 , S 2 , K 1 , and O 1 constituents sampled from the He and Weisberg [2002] WFS tide model, and a radiation p ffiffiffiffiffi boundary condition with a propagation speed of gh is used for velocity. Temperature is held constant throughout, and a nongradient open boundary condition is applied to salinity such that outgoing water may leave with its prognostic salinity, whereas incoming water enters with a salinity of 35 psu. To avoid instability the tidal amplitudes are ramped up from zero to their full values over the first M 2 tide cycle. [ 13 ] Three different classes of experiments are conducted. The first is a barotropic tide simulation. The second includes tides and rivers with prognostic salinity. The third adds either constant or oscillatory winds to the tide, plus river case. [ 14 ] After the initial ramp up over one M 2 tide cycle the model is run to quasi-equilibrium over an additional 9 M 2 cycles. Sixty more M 2 cycles are then run for analysis. We employ a linear least squares harmonic analysis (following Foreman [1977, 1978]) to compute the amplitudes, phases, and current ellipse parameters for the M 2 , S 2 , K 1 , and O 1 constituents at each of the model grid points. The model- predicted tidal elevations agree well with observations at the Naples and Ft. Myers tide gauges (Figure 3). The model duplicates the spring to neap tide variation, including the more diurnal (semidiurnal) character at spring (neap) tide. The simulation at Naples is better than at Ft. Myers since Naples is closer to the open boundary where the tides are forced and Ft. Myers is affected by depth and geometry complexities. We found that by improving the bathymetric data from the Caloosahatchee River we improved the Ft. Myers simulation. Along with these sea level comparisons, Table 1 shows the observed and computed tidal current ellipse parameters at a mooring positioned southwest of Sanibel (see Figure 2). Combined, the sea level and velocity comparisons demonstrate the legitimacy of the open boundary condition in forcing the tides. [ 15 ] Model-predicted coamplitude and cophase charts for the M 2 , S 2 , K 1 , and O 1 tidal constituents are presented in Figures 4a and 4b, respectively. The local innershelf results are consistent with the shelf-wide findings of He and Weisberg [2002]. The semidiurnal species show amplitude gradients from north to south, whereas the diurnal species are spatially more uniform. The tides propagate across the innershelf and into the estuary through BGP, Captiva Pass, and the mouth of SCB. Through BGP, the M 2 , S 2 , K 1 , and O 1 constituent amplitudes first decrease by 2.5 cm, 2 cm, 1 cm, and 1 cm, respectively, before increasing farther north into the Peace River by 6 cm, 3 cm, 1.5 cm, and 2 ...

Similar publications

Article
Full-text available
Under the background of urbanization and climate change, flood disasters are getting more and more serious. Floods tend to shift from single site to multi-sites and from urban area to suburb area, however, the mechanism of flood formation and transformation at different scales under changing environment is still unclear. In this study, we proposed...
Article
Full-text available
In the Chilean region of Patagonian fjords and channels between the Penas Gulf and the western mouth of Magellan's strait, Evadne nordmanni, Podon leuckarti and Pseudoevadne tergestina cladocerans were identified. The analyzed organisms were captured with Bongo nets through oblique hauls between 200 m of depth to the surface, in 41 oceanographic st...
Article
Full-text available
The objective of this study was to evaluate the temporal distribution of the foraminifera microfauna of the Caravelas River estuary (Bahia) to determine the different stages of marine influence on this ecosystem. For this purpose, two core samples 60 cm in length were collected at different sites in the estuary. In the laboratory, subsamples were t...

Citations

... Located on the west coast of the Florida peninsula, the CRE and Charlotte Harbor are vital areas for human and ecological activities. The CRE is connected to the Charlotte Harbor Estuary and spans a total area of~800 km 2 , which makes it the second-largest estuarine system in Florida [19]. The CRE is 42 km in length, spanning from Franklin Lock and Dam (S-79) to Shell Point [20]. ...
Article
Full-text available
Residence time is an important parameter linked to the water quality in an estuary. In this paper, we identify and analyze the main processes that affect the residence time of the Caloosahatchee River Estuary, a micro-tidal and mixed diurnal-semidiurnal estuary located in western Florida. Multiyear validated hydrodynamic hindcast results were coupled with an offline particle tracking model to compute the residence time of the estuary, which showed a strong seasonality driven by the river discharge. The residence time reduced with increasing river flow. The wind velocity and direction also affected the residence time. The influence of the wind was dependent on the magnitude of the river discharge. In general, upstream-directed wind increased residence time, while downstream-directed wind decreased residence time. Downstream wind during the dry period reduced residence time on average by a week. Processes such as water density gradient-induced circulation and particle buoyancy also influenced the residence time of the estuary. The outcomes of this study can be used to better understand the influence of the main physical processes affecting the residence time at other similar estuaries and to help in the management of the estuaries to improve their water quality.
... The response of the harbor to Hurricane Irma was evaluated with the West Florida Coastal Ocean Model, which simulated the observed negative surge and corresponding drying of the estuary and Tampa Bay (Liu et al. 2020). The estuary and gulf circulation patterns were also assessed in a three-dimensional modeling study, which evaluated the influence of tides, rivers, and wind on estuarine hydrodynamics and salinity (Zheng and Weisberg 2004). These models extend from the tidal rivers to the Gulf of Mexico, providing information on regional water movement, without predicting hydrodynamic behavior in the narrow inland channels. ...
... https://doi.org/10.21428/f69f093e.4f085bec. Table 4 Heat map of percent change in surface water elevations between the proposed hydraulic reconnection and baseline configurations for each weather condition and analysis site provides a high-resolution assessment of water flow in a residential waterway as compared to large scale models of Charlotte Harbor, which yield regional results of circulation (Chen 2020;Kim et al. 2010;Liu et al. 2020;Zheng and Weisberg 2004). The simulated increase in tidal range has ecological implications, for tidal amplitude heavily influences the distribution and characteristics of estuarine vegetation (McLusky and Elliott 2004). ...
Article
Full-text available
Coastal environments around the globe are subject to anthropogenic stresses due to dense coastal populations. The response of development activities on dynamic estuarine ecosystems, influenced by tidal forces, freshwater flows, salinity variations, and intricate coastal land morphology, is often uncertain. This case study evaluates how connectivity and coastal geomorphology influence flow patterns by modeling the effects of a proposed hydraulic reconnection project on water movement between the Manchester Waterway, a coastal residential community, and Charlotte Harbor, a large open water estuary in the Gulf of Mexico. An unstructured grid, 2D model was developed utilizing Delft3D Flexible Mesh to simulate estuary hydrodynamics under proposed conditions for four different weather conditions, including recorded 2021–2022 weather, future sea level rise, an extreme weather event, and a combination of extreme weather and sea level rise. Simulated flow results for proposed conditions were compared to present day flow patterns for analysis of the predicted changes in water levels and velocity magnitudes in the waterway. The results show that increased connectivity between the Manchester Waterway and Charlotte Harbor is expected to increase tidal amplitudes largely due to a lowering of minimum water levels in the waterway. During storm events, water elevations are predicted to drop to lower elevations following peak storm surge due to proposed conditions, which may provide flooding relief. Model simulation results will aid hydraulic reconnection and guide a more comprehensive ecological restoration plan. This case study will also improve understanding of the major influencing forces in intricate estuarine environments and how ecosystems may respond to land development, sea level rise, and increasing magnitude and frequency of tropical storms.
... The Caloosahatchee River Estuary is connected to the Charlotte Harbor Estuary through Matlacha Pass and Pine Island Sound. Zheng and Weisberg (2004) conducted a numerical study on the circulation features within the Charlotte Harbor Estuary, located north of the Caloosahatchee River Estuary. Their study concluded that the flow and salinity fields respond to a combination of tide, rivers, and wind. ...
... Tides in the Caloosahatchee River Estuary are mixed-semidiurnal dominant, with major constituents of M2, K1, O1, and S2 (Goodwin and Michaelis, 1976;Zheng and Weisberg, 2004). To evaluate our model performance, simulated tides are compared at 14 locations with predictions from NOAA (black dots in Fig. 1). ...
... This approach prevents model instability when a strong wind stress is applied over shallow depths. Note that wind-driven surface flow is commonly modeled in shallow coastal marshes (e.g., [31,32]), but wind effects are excluded in the DW approximation. In Frehd and Frehg, Equations (3) and (4) are solved with a semi-implicit, hybrid finite difference/finite volume method that is similar to Casulli [33], Casulli and Cattani [34]. ...
Article
Full-text available
A new high-performance numerical model (Frehg) is developed to simulate water flow in shallow coastal wetlands. Frehg solves the 2D depth-integrated, hydrostatic, Navier–Stokes equations (i.e., shallow-water equations) in the surface domain and the 3D variably-saturated Richards equation in the subsurface domain. The two domains are asynchronously coupled to model surface-subsurface exchange. The Frehg model is applied to evaluate model sensitivity to a variety of simplifications that are commonly adopted for shallow wetland models, especially the use of the diffusive wave approximation in place of the traditional Saint-Venant equations for surface flow. The results suggest that a dynamic model for momentum is preferred over diffusive wave model for shallow coastal wetlands and marshes because the latter fails to capture flow unsteadiness. Under the combined effects of evaporation and wetting/drying, using diffusive wave model leads to discrepancies in modeled surface-subsurface exchange flux in the intertidal zone where strong exchange processes occur. It indicates shallow wetland models should be built with (i) dynamic surface flow equations that capture the timing of inundation, (ii) complex topographic features that render accurate spatial extent of inundation, and (iii) variably-saturated subsurface flow solver that is capable of modeling moisture change in the subsurface due to evaporation and infiltration.
... Three-dimensional numerical circulation models (e.g., Fringer et al., 2006;Hervouet, 2000;Luo et al., 2013), which now are applied routinely to embayments (Weisberg & Zheng, 2006), banktops (Purkis et al., 2017(Purkis et al., , 2019, and estuaries (Zheng & Weisberg, 2004), are a powerful tool for understanding water circulation in shallow areas. However, these models typically require bathymetric, wind, and tide data for model initialization and forcing. ...
Article
Full-text available
Abstract In shallow coastal regions, tides often control the water flux, which in turn directs sediment transport, nutrient delivery, and geochemical gradients. However, tides in shallow areas are spatially heterogeneous, making it challenging to constrain the geographic structure of tidal phase and amplitude without extensive networks of tide gauges. We present a simple remote sensing method for deriving tidal structure from satellite time series. Our method is based on two observations: (1) Tidally driven variations in water depth can be detected as changes in pixel intensity in optical satellite imagery, and (2) repeating passes by an orbiting satellite capture a region at different phases of the tidal cycle. By stacking multiple satellite acquisitions of a shallow bank, we can compute the relative tidal phase and amplitude for each pixel location, thereby resolving a detailed map of tidal propagation and attenuation. While our method requires a set of local water‐depth measurements to calibrate the color‐to‐depth relationship and compute tidal amplitude (in meters), our method can produce spatial estimates of tidal phase and relative amplitude without any site‐specific calibration data. As an illustration of the method, we use Landsat imagery to derive the spatial structure of tides on the Great Bahama Bank, estimating tidal phase and amplitude with mean absolute errors of 15 min and 0.15 m, respectively.
... Usually, during periods dominated by strong currents, the flow in the straits is unidirectional. Similar to two-layer flow driven by gravitational convection in tidal estuaries (Zheng and Weisberg, 2004), currents in straits are also two-directional, given that the vertically integrated transport decreases below a critical value (Stanev et al., 2017. Thus, the different secondary-circulation regimes in the NBTZ straits are expected to be similar to those in tidal estuaries, and this expectation will be scrutinized in the present study. ...
Article
Full-text available
In this paper, we explore the secondary flows in the Danish Straits using observations and numerical simulations performed with the unstructured-grid hydrodynamic model SCHISM covering the North Sea and Baltic Sea. The straits are resolved on scales of up to ∼100 m. Given that large-scale atmospheric variability dominates the transport in these straits, we focus on the processes with subtidal time scales. Similarities and differences between the in- and outflows in the straits and flood and ebb flows in estuaries are analyzed. Contrary to the tidal straining in estuaries, the Danish Straits feature substantial differences in the stratification stability during the outflow and inflow phases. With a resolution of ∼100 m, new transport and mixing pathways that were previously unresolved appear fundamental to the strait dynamics. The variety of the strait morphology leads to high variability in the appearance of secondary circulation. Helical cells, often with a horizontal extension of ∼1 km, develop in the deep parts of the channels. A comparison between the high-resolution simulation and a simulation with a coarse grid of ∼500 m in the straits suggests that the coarser resolution overestimates the stratification and misrepresents the transport balance; the axial velocities and transport through the Sound are underestimated by ∼12%. These differences are explained by the missing secondary circulation when the coarse resolution is used (approximately two grid-points per cell instead of ten grid-points per cell in the fine resolution model), along with the resulting changes in mixing along the straits. In conclusion, the use of ultrafine resolution grids is essential to adequately resolve secondary flow patterns and two-layer exchange. Thus, the problems caused by the failure to resolve the secondary circulation in straits appear similar to the problems caused by the failure to resolve mesoscale eddies in ocean models.
... Additionally, the region is regarded as being subtropical because of distinct wet (May-October) and dry (November-April) seasons (Sun, Wan, and Qui, 2016) with average total seasonal precipitation of 780.9 and 320.7 (mm), respectively (Dye, 2018). The estuarine system has a mean depth of 2-3 m and covers a surface area of~700-800 km 2 (Harris et al., 1983;Poulakis et al., 2004;Zheng and Weisberg, 2004). Marine water from the Gulf of Mexico enters the estuarine system through various inlets (e.g., Boca Grande Pass, Captiva Pass, Redfish Pass, and Big Carlos Pass) between the barrier islands and San Carlos Bay ( Figure 1). ...
... MIKE's flexible mesh (FM) technology (DHI, 2016) was used to construct the computational mesh, comprising 20328 nodes and 33875 triangular elements (Figure 2A,B). The model domain encompasses Gasparilla Sound, Charlotte Harbor, Pine Island Sound, Matlacha Pass, the CRE, San Carlos Bay, Estero Bay, and all of the major tributaries (see Figure 1) and extends 80 km offshore into the Gulf of Mexico to allow tides to rhythmically propagate from the outer shelf into the shallow inlets and bays (Figure 2A,B; Zheng and Weisberg, 2004). The FM is categorized into five distinct regions, depending on the preferred level of resolution. ...
... Previous modeling studies by Goodwin (1996) and Zheng and Weisberg (2004) found the estuarine system to be slightly ebb dominate (i.e. discharge is greater during ebb tidal phase) but with less ebb dominance throughout neap tides. ...
Article
Dye, B.; Jose, F., and Allahdadi, M.N., 0000. Circulation dynamics and seasonal variability for the Charlotte Harbor Estuary, Southwest Florida coast. Journal of Coastal Research, 00(0), 000-000. Coconut Creek (Florida), ISSN 0749-0208. A hydrodynamic model was developed and validated for the Charlotte Harbor estuarine system, located in SW Florida, to elucidate freshwater fluxes within the system's various inlets during diverse hydrologic conditions. Fresh water entering the system not only varies seasonally but also, because of regulatory fresh water, releases controlling water levels within an upstream lake. The unnatural freshwater releases have been found to negatively affect the system's ecology, in particular within the Caloosahatchee River portion of the system. Neither the flood nor ebb phase exhibits uniform dominance in flushing the system's four major passes. Boca Grande Pass and Big Carlos Pass were mostly ebb dominant, whereas San Carlos Bay was largely flood dominant; neither phase dominated at Captiva Pass. The similarities and/or contradictions of these results in comparison to former field and modeling results are mainly attributed to the differences between the freshwater sources and environmental forces corresponding to each study that forces a different mass-balance condition over the estuary-bay system and, thereby, at each individual inlet. A Lagrangian particle tracking study revealed particles released within the Peace River during different hydrological conditions were comparably transported regardless of freshwater inputs and predominate wind direction. In contrast, particles released within the Caloosahatchee River were flushed into the Gulf of Mexico within 10 days during a usually wet El NiñoNi˜Niño, dry (November-April) season period whereas during the summer wet (May-October) season released particles remained in the estuary for a longer period (13 days), ultimately resulting in their further transport into Pine Island Sound and Matlacha Pass. The results also demonstrate the effect of freshwater river inputs and wind on the travel time of the neutrally buoyant particles within the estuarine system. The hydrodynamic and coupled particle tracking model serve as the first step in a forthcoming larval transport modeling study. ADDITIONAL INDEX WORDS: Hydrodynamic model, Caloosahatchee River, shallow estuary, MIKE model, tidal inlets, particle tracking.
... Additionally, the region is regarded as being subtropical because of distinct wet (May-October) and dry (November-April) seasons (Sun, Wan, and Qui, 2016) with average total seasonal precipitation of 780.9 and 320.7 (mm), respectively (Dye, 2018). The estuarine system has a mean depth of 2-3 m and covers a surface area of~700-800 km 2 (Harris et al., 1983;Poulakis et al., 2004;Zheng and Weisberg, 2004). Marine water from the Gulf of Mexico enters the estuarine system through various inlets (e.g., Boca Grande Pass, Captiva Pass, Redfish Pass, and Big Carlos Pass) between the barrier islands and San Carlos Bay ( Figure 1). ...
... MIKE's flexible mesh (FM) technology (DHI, 2016) was used to construct the computational mesh, comprising 20328 nodes and 33875 triangular elements (Figure 2A,B). The model domain encompasses Gasparilla Sound, Charlotte Harbor, Pine Island Sound, Matlacha Pass, the CRE, San Carlos Bay, Estero Bay, and all of the major tributaries (see Figure 1) and extends 80 km offshore into the Gulf of Mexico to allow tides to rhythmically propagate from the outer shelf into the shallow inlets and bays (Figure 2A,B; Zheng and Weisberg, 2004). The FM is categorized into five distinct regions, depending on the preferred level of resolution. ...
... Previous modeling studies by Goodwin (1996) and Zheng and Weisberg (2004) found the estuarine system to be slightly ebb dominate (i.e. discharge is greater during ebb tidal phase) but with less ebb dominance throughout neap tides. ...
... Tidal currents in many estuaries are complex spatially, and vary temporally. Accurate representation of these currents often requires the use of numerical simulations [30][31][32][33]. ...
... The differences in the results are the consequence of the added forcing and its nonlinear interaction with the other forcing. This procedure has been followed by several authors (Bolaños et al., 2013;Li and Li, 2012;Souto et al., 2003;Weisberg and Zheng, 2006;Zheng and Weisberg, 2004), to investigate the effects of tides, wind, and river discharge over simulation times varying from weeks to a few months. Although the effect of up-and downestuary winds and up-and downwelling favorable winds granted a seasonal perspective to some investigations (Li and Li, 2012;Weisberg and Zheng, 2006;Zheng and Weisberg, 2004), there was no effort to simulate the seasonal cycle. ...
... This procedure has been followed by several authors (Bolaños et al., 2013;Li and Li, 2012;Souto et al., 2003;Weisberg and Zheng, 2006;Zheng and Weisberg, 2004), to investigate the effects of tides, wind, and river discharge over simulation times varying from weeks to a few months. Although the effect of up-and downestuary winds and up-and downwelling favorable winds granted a seasonal perspective to some investigations (Li and Li, 2012;Weisberg and Zheng, 2006;Zheng and Weisberg, 2004), there was no effort to simulate the seasonal cycle. An exception is the work of Teixeira (2010), who investigated the seasonal variation in the advection of salt and heat by the density-and wind-driven circulation of the temperate, hypersaline Spencer Gulf in South Australia. ...
Article
Santana, R.; Teixeira, C., and Lessa, G., 2018. The impact of different forcing agents on the residual circulation in a tropical estuary (Baía de Todos os Santos, Brazil). This study investigated the forcing of the residual circulation in a tropical estuary, Baía de Todos os Santos, Brazil (BTS). A numerical model (Regional Ocean Modeling System), forced with climatological means, was used to investigate the seasonal roles of the tide, wind, net heat, surface water fluxes, and river discharge on the residual circulation. The tide is the main driver of the circulation and its residual flow is structured at the bay mouth with a net-landward flow in the channel center and a net-seaward flow on the shoulders. The addition of wind drag affected the surface circulation, forcing W-bound (N-bound) currents in the spring-summer (fall-winter) months and generating a positive sea-level slope toward the continent. The addition of heat and surface water fluxes to the simulation established an incipient gravitational circulation, mainly because of the large (approximately 0.2°C/km) temperature gradients in the summer. The river discharge, added last, was second to the tide in driving the residual circulation in the BTS and established the gravitational circulation through most of the bay, including the bay mouth. The residual velocities attained magnitudes smaller than 0.13 ms⁻¹ in the main channel and were seasonally controlled, with stronger currents during the summer when the river discharge peaked. The onset of the baroclinic forcing (heat and surface water fluxes plus river discharge) reduced the e-folding flushing time by a factor of five relative to the experiment with barotropic forcing (wind and tides) only.