You are on page 1of 15

Journal of Marine Systems 236 (2022) 103804

Contents lists available at ScienceDirect

Journal of Marine Systems


journal homepage: www.elsevier.com/locate/jmarsys

On the mechanisms controlling near-coast circulation in the southern


Colombian Pacific at tidal, seasonal, and interannual time scales
Óscar Álvarez-Silva a, *, Víctor Saavedra b, c, Luis Otero a, Juan C. Restrepo a
a
Department of Physics and Geosciences, Universidad del Norte, 080007 Barranquilla, Colombia
b
Universidad Católica Luis Amigo, Facultad de Ingenierías y Arquitectura, Medellín, Colombia
c
OCEANICOS Research Group, Universidad Nacional de Colombia, Medellín, Colombia

A R T I C L E I N F O A B S T R A C T

Keywords: This research analyzed the near-coastal circulation along 300 km in the southern Colombian Pacific, a meso-tidal
Coastal circulation region formed by Tumaco Bay and the Mira River Delta. The interaction between tidal dynamics and the Mira
Residual flow River plume stratification and dispersion is not well known. Moreover, the combined effects of tide and density
Tidal straining
gradients on the circulation patterns in Tumaco Bay have rarely been studied. The region was investigated using
Mira River
Tumaco Bay
the Delft3D hydrodynamic model, calibrated and validated using field data. The results show that Tumaco Bay is
ebb-dominated, and the tide and bottom shape are the dominant forces in the circulation patterns inside the bay.
The weak horizontal density gradient induced by Mira River did not show a substantial effect on the circulation
of the bay. Near the mouth of the Mira Delta, the water column was predominantly partially mixed, but strat­
ification changed within the tidal cycle, presenting tidal straining along the minor axis of the plume. This may
generate strongly stratified conditions in the water column during flood periods. Although tidal straining is very
common along the major axis of tide-dominated estuaries, it is not as common in coastal river plumes. The
analyzed system provides additional evidence about this phenomenon in a tropical delta.

1. Introduction Circulation patterns in semi-enclosed water bodies such as Tumaco


Bay determine water residence times, the accumulation or dispersion of
River mouths and semi-closed water bodies are strongly dynamic pollutants and nutrients, and trends toward erosion or accumulation of
coastal systems controlled by complex oceanographic, fluvial, geomor­ sediments in those water bodies (Murphy and Valle-Levinson, 2008;
phological, and climatic processes (Kjerfve, 1979, 1990). The interplay Yang et al., 2004). Circulation within bays is determined by their
of these dynamics leads to the formation of patterns of circulation, morphology, density gradients, tides, and winds. Frequently, in meso-
stratification/mixing, sediment transport, erosion/sedimentation, and and macro-tidal zones such as Tumaco Bay, the tide is the primary
morphological evolution that are unique to each system (Dyer, 1997). forcing of the system (Alosairi et al., 2018; Sigaúque et al., 2021).
Understanding the interrelation between forcings and their effects in Meanwhile, the discharge of freshwater from the Mira river into the sea
coastal zones facilitates the protection of strategic ecosystems, the generates a region of fluvial influence (ROFI) where density gradients
reduction of vulnerability to natural threats, the sustainable use of become an important factor in the coastal hydrodynamics (Simpson,
natural resources, and, in general, the integrated management of coastal 1997) and the ecological balance of the area (Horner-Devine et al.,
zones (LOICZ, 2005). 2009). Even more interesting is the interaction between the two forc­
The southern Colombian Pacific coast embraces the Bay of Tumaco ings, the river plume and tide in areas where both are dominant (Chao,
and the mouth of Mira River (Fig. 1). This is a region of exceptionally 1990; Guo and Valle-Levinson, 2007; Jay et al., 2009).
high natural value and broad biodiversity. It also has large ecotourism Despite the importance of hydrodynamics to identify its effects on
and agro-industrial potential. The study area presents a mixed behavior the conservation, navigability, development, and economy of the region;
in which the tide and river discharge are important forcings of coastal to the best of the authors' knowledge, there is no peer-reviewed research
processes. In addition, a strong temporal variability associated with describing the circulation in Tumaco Bay and Mira River plume, and
seasonal and interannual changes in environmental forcings is expected. analyzing the interaction between both features. The present study

* Corresponding author.
E-mail address: oalvarezs@uninorte.edu.co (Ó. Álvarez-Silva).

https://doi.org/10.1016/j.jmarsys.2022.103804
Received 14 April 2022; Received in revised form 2 August 2022; Accepted 24 August 2022
Available online 31 August 2022
0924-7963/© 2022 Elsevier B.V. All rights reserved.
Ó. Álvarez-Silva et al. Journal of Marine Systems 236 (2022) 103804

flows, and mean sea level, among others (Devis-Morales, 2003a, 2003b;
Restrepo and Lopez, 2007).

3. Materials and methods

3.1. Field data

Data used for characterization of study zone and model forcing,


calibration, and validation is taken from secondary sources. Water level
data from 1952 to 2014 (with 15% gaps) was obtained from the Sea
Level Station Monitoring Facility by UNESCO (http://www.ioc-sea
levelmonitoring.org/station.php?code=tumc2). This data was recor­
ded hourly at location P1 in Fig. 1. Statistical distribution of the tidal
range according to these data is shown in Fig. 2A,B. Astronomical tides
strongly dominate the water level here. An analysis of the meteorolog­
ical residual of water level after extracting the astronomical components
shows that the average range of this residual is 0.21 m, equivalent to
8.5% of the mean tidal range, and only 9.7% of the time, the meteoro­
logical tide is larger than 0.50 m (equivalent to about 20% of the mean
tidal range).
Meanwhile, wind speed and direction were obtained from a meteo­
rological station by the General Maritime Directorate of Colombia
(DIMAR), also located at P1. These data extend from 2011 to 2019 with
Fig. 1. Location and description of the study area. The modelling domain is an hourly resolution and 6% of gaps. The statistical distribution and
enclosed. Sea level and wind measurements are available at P1. Water current climatology of data are shown in Figs. 2E-H and 3A, respectively.
measurements are available at P2. Green lines delimit the domain of the nested Finally, daily river discharge was recorded by the Institute of Hydrology,
grids in the modelling. Red transects indicate locations where model results Meteorology and Environmental Studies (IDEAM) from 2000 to 2017,
were analyzed in detail. (For interpretation of the references to colour in this with 9% gaps, at a limnometric station located 75 km upstream of the
figure legend, the reader is referred to the web version of this article.)
river mouth. Statistical distribution and climatology of these data are
shown in Figs. 2C-D and 3B, respectively.
addresses this issue for different representative scenarios of the seasonal Water velocity measurements in the region are scarce. Reliable
and interannual climatic variability in the study area. To do this, the measurements in the water column are available only for 32 h from April
Delft3D hydrodynamic model was implemented and validated as 14th – 15th, 2010. This time series was recorded at location P2 in Fig. 1
described below. by DIMAR, a location of about 10 m depth from mean sea level. These
measurements were made throughout the entire water column every
2. Study zone 1.0 m, using an ADCP located at the surface looking downward with a
blanking distance of 1.5 m. Temporal resolution between samples was
The southern Colombian Pacific is a meso-tidal zone with weak 10 min.
winds. According to the tide gauge at point P1 in Fig. 1, the area has a
semidiurnal tidal regime with a range of 2.47 m ± 0.61 m and 10th and 3.2. Hydrodynamic modelling
90th percentiles of 1.67 and 3.32 m, respectively (Fig. 2A-B). The winds
at P1 show weak speeds with an average of 2.53 m/s (Fig. 2E-F). The The Delft3D three-dimensional hydrodynamic model was applied to
main directions are from the west, southwest, and south, which sum up analyze the study area's circulation patterns. Delft3D simulates two-
79% of the probability of occurrence (Fig. 2G-H). dimensional (2DH, vertically averaged) or three-dimensional (3D)
Tumaco Bay has an area of 445 km2 and a semi-round shape, with the flow and transport phenomena resulting from forcing by baroclinic and
principal axis 24 km long and a width at the mouth of 21.3 km. Depths in barotropic pressure gradients, stresses induced by wind and atmospheric
the bay vary between 2 and 35 m. At the southwestern extreme of the pressure gradients, friction at the bottom and lateral edges, earth rota­
bay is the largest population in the Colombian Pacific, with 138,000 tion (Coriolis force), rivers flows, surface heat exchange, and internal
inhabitants in 2018 (Dane, 2018). Here is located the second most waves, among others (Lesser et al., 2004; Deltares, 2021).
important seaport on the Colombian Pacific coast. To the south of the Delft3D-FLOW solves the shallow water equations in 2D and 3D. This
bay is the mouth of the Mira River. This river has a discharge of 790 ± system of equations includes the continuity equation, horizontal mo­
456 m3/s with 10th and 90th percentiles of 297 m3/s and 1413 m3/s, mentum equations, and transport equations for conservative constitu­
respectively (Fig. 2C-D). Its sediment discharge is estimated at 9.7 × 10 6 ents. The equations are formulated in planar or spherical orthogonal
t/yr (Restrepo and Lopez, 2007), making it one of the most important curvilinear coordinates. Spatial discretization is performed via finite
tributaries on the Pacific coast of South America. differences with curvilinear, orthogonal, and structured grids. The
To analyze the climate variability in more detail, Fig. 3 presents the model offers two vertical coordinate systems: the so-called σ coordinate
monthly mean annual cycle for the freshwater discharge of Mira River (σ model) and Z Cartesian coordinate (Z model) systems.
and winds over Tumaco Bay (P1). Fig. 3 shows that the seasonal regime The model uses the Boussinesq approximation for shallow waters to
of flows and winds are in counterphase. The largest average flows and resolve the vertical velocities. Additionally, the model solves the equa­
weakest winds occur between March and May (wet season). On the tion of state and transport of conservative constituents such as salinity
contrary, the lowest flows occur between July and September (dry and heat and includes various closure models to resolve turbulence. This
season), when the winds are the strongest (Devis-Morales, 2003a; model has been widely applied to coastal and estuarine areas (Alosairi
Restrepo, 2006). et al., 2018; Escobar et al., 2015; Hu et al., 2009; Sigaúque et al., 2021;
Previous studies of the Colombian Pacific coast have determined that Zarzuelo et al., 2017; Zhu et al., 2017).
the El Niño / Southern Oscillation (ENSO) is an important forcing in the
hydrology and oceanography of the area, influencing precipitation rates,

2
Ó. Álvarez-Silva et al. Journal of Marine Systems 236 (2022) 103804

20%
8000 1

A 18% 17%
0.8
B
No. observations

6000
13%
11% 0.6
4000
7% 8% 0.4

2000 4%
0.2
1%
1%
0 0
1 1.5 2 2.5 3 3.5 4 1 1.5 2 2.5 3 3.5 4
Tidal range (m) Tidal range (m)

C D

E F

G H

Fig. 2. Statistics of environmental forcings in the study area: A, B) tidal range; C, D) Mira River flow; E, F) wind speed; G, H) wind direction. A, C, E, G) probability
distribution histogram with numbers above bars indicating the probability of occurrence of each interval. B, D, F, H) cumulative probability function.

3.2.1. Model configuration cells of constant size. The results of a coarser grid covering a larger area
The simulated coastal and maritime area has a coastline of 278 km are used to define the boundary conditions of the next finer grid, which
and an area of ~6225 km2, centered on the municipality of Tumaco. The comprises a subdomain of the larger grid. The first general modelling
major axis of the domain is oriented northeast-southwest, coinciding grid (M1, Fig. 1) has a horizontal spatial resolution of 500 m × 500 m. It
with the alignment of the coast and bathymetry (Fig. 1). The imple­ covers the Mira River Delta and Tumaco Bay. A second intermediate
mented model has three offline nested structured grids, all with square transition grid (M2) with a spatial resolution of 150 m × 150 m is used as

3
Ó. Álvarez-Silva et al. Journal of Marine Systems 236 (2022) 103804

Fig. 3. Boxplot of the climatology of A) freshwater discharge of Mira River and B) winds in the study area. The blue line shows multi-year monthly averages, and the
red lines standard deviation. Thick bars represent the range of variability between the 25th and 75th percentiles, and thin bars represent three standard deviations.
Circles show data beyond three standard deviations. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of
this article.)

a transitional step to avoid numerical instabilities in nesting due to large bathymetric information was provided by the General Maritime Direc­
cell-size transitions (Pham et al., 2016; Vannitsem and Chomé, 2005; torate of Colombia. These data have a variable spatial resolution of up to
Hodges, 2013), and a third detailed grid (M3) with 50 m × 50 m reso­ 20 m × 20 m and cover 5625 km2 of the 10,000 km2 simulation domain.
lution that covers the Tumaco Bay (Fig. 1). For the rest of that domain, bathymetric data were taken from the
A Sigma-type grid was implemented for vertical discretization using GEBCO database (General Bathymetric Chart of the Oceans, www.
15 vertical cells of variable width with more refinement near the surface gebco.net), which has global coverage and 15′′ arc resolution.
and bottom. A 3D model configuration allows analyzing the vertical
circulation and mixing processes associated with the Mira River plume 3.2.2. Boundary conditions
in the coastal zone. The K-E turbulent closure model was used, as well as For the general grid M1, level and salinity conditions were imposed
the cyclic advection scheme for momentum. The computational time on the open boundaries with the Pacific Ocean, allowing the model to
step of each grid to achieve numerical stability and convergence -after resolve velocities. On the other hand, boundary conditions of water
several testing- was 15 s for grids M1 and M2 and 12 s for grid M3. levels, salinities, and velocities on grids M2 and M3 resulted from sim­
For Tumaco bay and coastal areas of the Mira River Delta, the ulations of the immediately previous coarser grid.

4
Ó. Álvarez-Silva et al. Journal of Marine Systems 236 (2022) 103804

Table 1
Representative scenarios of seasonal and interannual climatic variability of the study area selected for modelling.
Interannual variability Seasonal variability Scenario number Month Mean river discharge [m3/s] Mean wind speed [m/s] Mean wind direction [◦ ]

Wet season 1 April 2012 756 ± 91 2.4 ± 1.2 230 ± 91 ◦


No ENSO
Dry season 2 August 2012 491 ± 82 2.9 ± 0.9 253 ± 44 ◦
Wet season 3 April 1998 1181 ± 179 3.6 ± 1.4 226 ± 84 ◦
El Niño
Dry season 4 August 1997 399 ± 95 4.4 ± 1.8 212 ± 37 ◦
Wet season 5 April 2011 1226 ± 189 2.4 ± 1.0 247 ± 73 ◦
La Niña
Dry season 6 August 2010 472 ± 158 5.5 ± 1.5 210 ± 23 ◦

Sea level at the boundaries of grid M1 includes the four dominant phases were taken from NASA's TPXO 9 model (https://www.tpxo.net/
semidiurnal harmonics of the astronomical tide (M2, S2, N2, and K2), global). As shown in Section 3.1, water level is well represented by as­
the four main diurnal harmonics (K1, O1, P1, and Q1), and the three tronomical tide, given the relatively low amplitudes of meteorological
main long-period harmonics (MF, MM, and SSA). The amplitudes and oscillations and the scarcity of large events. The salinity boundary

Fig. 4. A, C) Comparison between measured (red) and simulated (black) sea level in calibration period. B, D) Quantification statistics of the goodness of results. A, B)
Results for grid M1 and C, D) for grid M2. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

5
Ó. Álvarez-Silva et al. Journal of Marine Systems 236 (2022) 103804

Fig. 5. A, C) Comparison between measured (red) and simulated (black) sea levels in validation period. B, D) Quantification statistics of the goodness of results. A, B)
Results for grid M1 and C, D) for grid M2. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

conditions of the general grid were defined as constant in time and space current profile at point P2 (Fig. 1), for which data are available for 32 h
with a value of 35 g/kg, consistent with oceanic conditions in the area between April 14–15, 2010.
far from the coast, according to the HYCOM database. Water tempera­ The calibration parameters for the M1 grid were the vertical eddy
ture was not simulated because it has low importance in density viscosity coefficient of the turbulent closure model Κ - Σ (Kappa-Epsilon)
compared to salinity in coastal environments with fluvial influence. and the numerical boundary reflection parameter -for a larger reflection
For the three grids, uniform winds were used throughout the domain, parameter, less reflective is the open boundary for short wave distur­
corresponding to data from the meteorological station at P1 for the bances that propagate toward the boundary from inside the model
modeled periods. For modelling prior to 2011, when measured wind (Deltares, 2021)-. For grids M2 and M3, which were in shallower areas,
data were unavailable, data from the CFSR reanalysis model (Saha et al., the most critical calibration coefficient was the bottom friction (Mann­
2014) were used. Empirical distribution functions of wind speed for both ing) coefficient. Wind drag coefficient was not calibrated, and default
datasets were compared to ensure similarity, showing a good agreement. values were used.
Finally, the fluvial discharge of Mira River was represented using mean
daily flows. This data was smoothed by a 7-day moving average to fill 3.2.4. Modelling scenarios
the data gaps on the data observed during some simulation scenarios. Circulation patterns in the study area were analyzed from repre­
sentative scenarios of the seasonal and interannual climatic variability
3.2.3. Calibration and validation of the natural system. To consider seasonal climatic variability, the
The modeled period in the calibration stage was April 2012 because monthly variability of freshwater cycle of Mira River and winds over
2012 is considered an average climatological year, and April is repre­ Tumaco Bay shown in Fig. 3 were analyzed. These two factors are
sentative of the rainy season when high river discharges take place. The considered because the river discharge influences water density, circu­
modeled period in the validation stage was August 2012 because this lation, and mixing. The winds affect the dispersion of the river plume in
period is representative of the climatic dry season when low river dis­ the coastal zone. As mentioned before, river flow and winds are un
charges take place. Calibration and validation are made for opposite counterphase. Accordingly, April and August were selected to represent
climatological scenarios. This permitted the verification of satisfactory the wet and dry seasons, respectively, which coincide with the extremes
model performance under different conditions. of seasonal variability.
For calibration and validation of grids M1 and M2, field data and On the other hand, the main interannual climate forcing at the global
model results of sea level were compared. Because of the meso-tidal level, mainly in the tropical Pacific Ocean, is the ENSO macroclimatic
character of the study area, variation of the free surface is a dominant phenomenon (Cai et al., 2020; Poveda et al., 2001). According to the
forcing in the dynamics of the area, and its adequate representation is Physical Science Laboratory of the North American Oceanic and Atmo­
determinant for reproducing the circulation patterns. The M3 grid was spheric Administration, the strongest positive ENSO anomalies in the
validated by comparing measurements and modelling results of the last 40 years occurred in the 1997–1998 El Niño event (Wolter and

6
Ó. Álvarez-Silva et al. Journal of Marine Systems 236 (2022) 103804

days. For the detailed grids, the Manning coefficient with the best results
was n = 0.04.
A Fig. 4 compares measured and simulated sea levels at point P1 for
grids M1 and M2 with the configuration described above. The model
adequately captured both the amplitude and phase of the tide, including
the transition between spring and neap tides. However, in grid M1
(Fig. 4A), the model underestimated the amplitude by an average of
~0.3 m during low waters in spring tides. On the contrary, on the nested
grid M2 (Fig. 4C), the amplitude of the tidal wave was overestimated by
~0.3 m. The difference in the amplitude of the tidal wave between grids
is attributed to changes in the bay's total volume when changing the
model's resolution from one grid to another, especially in narrow
channels like that where the P1 sensor is located. However, Fig. 4B,D,
the similarity between measured and simulated data is confirmed by a Q
– Q plot and the determination coefficient R2, which had satisfactory
value of 0.92 for grid M1 and 0.91 for grid M2.
Additionally, Fig. 5 compares model results and field measurements
B during the validation. The upper panels (Fig. 5A-B) correspond to the
validation of the M1 grid and the lower panels (Fig. 5C-D) to the M2
grid. The left column (Fig. 5A,C) compares sea levels measured. Again,
the model adequately captured both the amplitude and phase of the tide.
The right column (Fig. 5B,D) confirms the similarity between the
measured and simulated data via a Q – Q plot, obtaining determination
coefficients R2 = 0.94 for the M1 grid and R2 = 0.95 for the M2 grid.
These values are even larger than for the calibration period, and the
RMSE is smaller for both grids.
For the general grid M1, there was a smaller amplitude in the model
results than in field data, as in calibration. For grid M2, the opposite
occurred, i.e., the model yielded amplitudes greater than the real ones.
However, we believe that in general, the model response was
satisfactory.
To validate modelling of the M3 grid, we used measurements of
Fig. 6. Validation of measured and simulated velocities for M3 grid. Measured velocity magnitude and direction at point P2 in Fig. 1 for April 14 and
and simulated currents at 5 m depth at point P2, shown in Fig. 1. 15, 2010. These measurements were compared with modelling results
for the same location and dates. Water velocity in both data and model
Timlin, 1998). Meanwhile, the strongest negative ENSO anomalies in are very uniform along the water column in the observation point. Fig. 6
that period occurred during La Niña 2010–2011 (Hoyos et al., 2013). compares the eastward and northward components of the measured and
Therefore, these years were considered to represent the opposite ex­ simulated velocities, magnitude and direction at 5 m depth from mean
tremes of interannual climate variability. sea level. There is a good reproduction of the magnitude and direction of
Combining the seasonal and interannual climatic variability resulted flow velocities throughout the tidal cycle, except for underestimating
in the six scenarios presented in Table 1. It should be noted that ENSO the northward component, which does not considerably affect the
events usually start in the boreal summer of the first year and finish in magnitude or direction of the speeds. It is highlighted that the model
the boreal spring the following year. For this reason, for El Niño and La reproduced the cyclic changes of velocity direction in both magnitude
Niña events, the dry season scenario was taken as August of the first year and time.
of the phenomenon and the wet season as April of the second year.
The average pattern of salinity and circulation for each simulated 4.2. Dispersion of the Mira River plume
scenario was obtained from an average of 15 days of modelling, after 7
days of spin-up scenario. An average of 15 days was used because this The results in this and the following section are described simulta­
includes a complete spring and neap tides cycle. neously for the six evaluated scenarios. Each figure includes six panels,
where the left column shows the results of the three wet season scenarios
4. Results (1, 3, and 5) and the right column the dry season scenarios (2, 4, and 6).
The top row presents the results for non-ENSO years, the middle row for
This section starts with the results of the calibration and validation of El Niño years, and the bottom row for La Niña years.
the model. Afterward, surface patterns of circulation and salinity pat­ The mean salinity and surface residual velocity patterns for each
terns are presented, first for the M1 grid, focusing on the Mira River simulated scenario are presented in Fig. 7. There, the effect of the Mira
Delta; later, these same analyses are shown for the M3 grid with a focus River discharge on the general salinity pattern is predominant. The blue
on Tumaco bay. line represents a 5% concentration of river water (where salinity is 5%
lower than open ocean), which is used to delimit the maximum plume
extent. This river plume runs parallel to the coast without entering the
4.1. Model calibration and validation sea. The greatest differences in salinity and velocity are observed intra-
annually between climatic seasons and not interannually between ENSO
During the calibration of the model, the combination of parameters phases. This is because, as observed in Table 1, variations in the fresh­
that showed the best results was that with a turbulent viscosity coeffi­ water discharge of Mira River and wind speed and direction are more
cient of 1 × 100 m2/s, a turbulent diffusivity coefficient of 1 × 10− 4 m2/ significant between seasons than between ENSO phases in the repre­
s, and a boundary reflection parameter of α = 130,000 s2. Under this sentative simulated scenarios.
configuration, the simulations converged to a stable solution after five The plume extent is smaller toward the north in the wet season

7
Ó. Álvarez-Silva et al. Journal of Marine Systems 236 (2022) 103804

Fig. 7. Patterns of mean salinity and surface residual velocity and maximum extent of Mira River plume for all scenarios. A, C, E) Wet season. B, D, F) Dry season.
During A, B) non-ENSO years; C, D) El Nino; E, F) La Niña. Blue line shows the maximum plume extent. (For interpretation of the references to colour in this figure
legend, the reader is referred to the web version of this article.)

scenarios, even though river flows are greater. This is because, during influences all of Tumaco Bay (although with salinities >28 g/kg), while
the dry season, the wind speed is slightly higher than during the wet in the wet season, weaker surface residual currents restrict the plume
season, but, more importantly, the wind direction is more persistent. i.e., extent to the southwestern half of the bay. The Coriolis acceleration does
the standard deviation of wind direction is lower, as shown in Table 1. not significantly affect the currents or plume dispersion because of the
Therefore, surface currents are stronger during the dry season with a equatorial location of the study area.
south-north direction near the coast, as seen by the arrows of the ve­ Very slow residual velocities were observed in Tumaco Bay, as usual
locity field in Figs. 7 B,D,F. Therefore, in the dry season, the river plume in areas where the tide dominates the circulation (Geyer and

8
Ó. Álvarez-Silva et al. Journal of Marine Systems 236 (2022) 103804

Fig. 8. Salinity difference between surface and bottom and maximum extent of Mira River plume for all scenarios. A, C, E) Wet season. B, D, F) Dry season. During A,
B) non-ENSO years; C, D) El Nino; E, F) La Niña. Blue line shows the maximum plume extent. (For interpretation of the references to colour in this figure legend, the
reader is referred to the web version of this article.)

9
Ó. Álvarez-Silva et al. Journal of Marine Systems 236 (2022) 103804

Fig. 9. Mean surface circulation (vectors) and salinity (blue lines) in Tumaco Bay for all scenarios. Bathymetry (thin black contours) and salinity difference between
surface and bottom (colored shades) are also shown A, C, E) Wet season. B, D, F) Dry season. During A, B) non-ENSO years; C, D) El Nino; E, F) La Niña. (For
interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

MacCready, 2014; Guo et al., 2018; Leonardi and Plater, 2017; Sigaúque Mira River Delta is observed, extending parallel but separated from the
et al., 2021). This tide produces cyclical currents on semidiurnal time coastline, reaching the southern end of Tumaco Bay. This stratification
scales that generate residual low-magnitude circulation patterns. varied between the wet and dry seasons, attaining a 20 g/kg difference
Fig. 8 depicts differences in salinity between the surface and bottom in salinity between the surface and bottom in the wet season. This is
due to the Mira River plume. Vertical salinity stratification in front of the because in the wet season, flows increase, and the plume extent

10
Ó. Álvarez-Silva et al. Journal of Marine Systems 236 (2022) 103804

This situation tends to export sediments and erode the bottom instead of
sedimenting the bay in the long term (Guo et al., 2018; Nidzieko, 2010;
Sigaúque et al., 2021).
The greatest magnitudes of the currents inside the bay are found in
the deepest zones, indicating that the flow is limited by bottom friction,
behaving as a hydrodynamically shallow system (Nidzieko, 2010; Valle-
Levinson, 2008). There are similar residual circulation patterns in all
scenarios, indicating that the tide and bathymetry dominate the study
area over the density gradients caused by the Mira River or the wind.
The same occurs in other meso- and macro-tidal bays slightly influenced
by fluvial discharges (Alvarez and Jones, 2002; Sigaúque et al., 2021;
Yang et al., 2004).
The fact that the residual currents are near-zero does not imply that
the instantaneous currents are also close to zero. Modelling Results (not
shown here) reveal that the instantaneous currents in spring tides can
reach 1.0 m/s during mid-ebb and mid-flood tides, albeit in opposite
directions.

5. Discussion

5.1. Circulation and stratification in Mira River plume

To observe the degree of stratification in the Mira River plume, its


Fig. 10. Fifteen-day averages of A, B) shoreward velocity; C, D) buoyant fre­
quency; E, F) shear stresses in the water column during A, C, D) wet season and
B, D, F) dry season, along transect T1 (Fig. 1). Positive velocities are toward
Tumaco Bay, i.e., out of the screen. All panels show isohalines every 2 g/kg.
Dashed line shows plume contour.

decreases, intensifying the vertical salinity gradients. On the other hand,


the vertical stratification rarely exceeded 10 g/kg in the dry season
during the dry season because of weaker river flows and a greater hor­
izontal dispersion of the plume.
Interannual changes were less noticeable than seasonal. However, in
the wet seasons of La Niña years and ENSO-neutral years, the vertical
difference of salinity in the coastal zone surrounding the mouth excee­
ded 25 g/kg, indicative of strong stratification.

4.3. Salinity structure and circulation in Tumaco Bay

Residual circulation patterns in Tumaco Bay, according to the more


detailed simulations of the M3 grid, are presented in Fig. 9. These pat­
terns are accompanied by bathymetry, using mean sea level as the
reference. The difference in salinity between the surface and bottom and
the surface isohalines are also shown. From the salinity patterns, it is
again seen that the Mira River influences the salinity structure of
Tumaco Bay, reducing the salinity of that bay horizontally. Likewise, it
is observed that the most remarkable differences in salinity occurred
seasonally, while in the same season, interannual variations were
negligible, again revealing the dominance of seasonality over ENSO in
determining the salinity structure.
The lowest salinities in the bay are found to the southeast, reaching
20 g/kg, while toward the north end of the bay, the salinity presents
oceanic values, with little influence from the Mira River. The effect on
the salinity of some minor drainages that discharge into the bay is also
observed. Vertically, the salinity is homogeneous because of mixing
induced by tidal currents and the vertical propagation of bottom tur­
bulence, which extended throughout the water column given the
shallow depths of the bay.
Meanwhile, residual currents inside Tumaco Bay are very weak,
barely exceeding 0.01 m/s in the deepest sectors of the bay entrance. As Fig. 11. Buoyant frequency and shoreline velocity (blue lines) along transect
mentioned above, this pattern of near-zero residual currents is charac­ T1 (Fig. 1) during a cycle of spring tides. Positive velocities are toward Tumaco
teristic of coastal systems dominated by cyclic currents generated by the Bay, i.e., out of the screen. All panels show isohalines every 2 g/kg. Dashed line
tide. In all scenarios, residual currents in the bay have a sustained shows plume outline. (For interpretation of the references to colour in this
seaward direction, indicating that the ebbing flow dominates the bay. figure legend, the reader is referred to the web version of this article.)

11
Ó. Álvarez-Silva et al. Journal of Marine Systems 236 (2022) 103804

relationship with circulation, and its variation with the tidal cycle, the
transect T1 shown in Fig. 1 was analyzed. Fig. 10A-B shows velocities
transverse to the transect, i.e., along the coast in section T1, for scenarios
1 (wet season) and 2 (dry season) according to Table 1. As described
above, the greatest variability in hydrodynamic patterns occurred
seasonally, so interannual cases were not considered in this analysis. The
T1 transect begins on the coast and extends seaward. Positive velocities
are toward Tumaco Bay (i.e., out of the screen). Also shown in the panels
are isohalines every 2 g/kg and a 5% concentration of river water (thick
dotted line), which delimits plume extent. Fig. 10C-D shows buoyant
frequency in the water column for both scenarios, estimated by N2 = −
(g/ρ0)(dρ/dz) (Stacey et al., 2012). This approximates the degree of
stratification of the water column. In strongly stratified conditions N2 >
0.01, whereas in vertically homogeneous conditions, N2 < 0.0025. In­
termediate values indicate that the water column is partially mixed
(Geyer et al., 2008). Finally, Fig. 10E-F presents the shear stresses pro­
duced by vertical changes in velocity. These were calculated by ∣τ∣ =
[( ) ( )2 ]1/2
2
km δu
δz + δv
δz (Ralston et al., 2010; Spicer et al., 2021), where
kmis the turbulent viscosity and u and v are the components of velocity
normal and parallel to the cross section.
As seen in Figs. 10A-B (and Figs. 7 and 9), velocities near the coast
have a northwesterly direction (toward Tumaco Bay), with a similar
pattern at both times of the year. However, speeds were faster in the wet Fig. 12. Fifteen-day averages of A, B) velocities perpendicular to the entrance
season, owing to a greater horizontal density gradient. On the other of Tumaco Bay (transect T2, Fig. 1); C, D) buoyant frequency and E, F) shear
hand, the plume presents a greater vertical extent in the dry season. stress in the water column during A, C, D) wet season and B, D, F) dry season.
This, added to the larger horizontal extent because of greater wind drag Positive velocities are seaward, i.e., toward the screen. All panels show isoha­
(Figs. 7 and 8), implies stronger dispersion of freshwater and, therefore, lines every 2 g/kg. Dashed line shows plume outline.
more mixed conditions in the vertical. The buoyant frequency shows
partially mixed conditions at both times of the year but slightly greater variations with the tidal cycle (Flores et al., 2020; Simpson and Souza,
stratification in the wet season, especially near the coast at depths <10 1995; Souza et al., 2008). This mechanism is unlikely to develop at low
m, as evident in Fig. 8 from the vertical salinity gradients. latitudes where the Coriolis acceleration is negligible, as is the case of
Meanwhile, shear stress is propagated from the bottom upward to the the Mira River. In contrast, with the Rhine, changes in stratification
river plume but dissipates at the interface between the two water appear to respond more to the successive contraction and expansion of
masses. Therefore, this stress is not contributing to plume mixing. Thus, the plume against the coast both vertically and horizontally as the tide
the mixing is produced more by tidal dispersion than by turbulence rises and falls, respectively.
associated with the propagation of shear stresses from the bottom. To the best of our knowledge, this observed mechanism of cross-axis
Fig. 11 shows the variability of circulation and stratification in a stratification variability in tropical river plumes has not been previously
cycle of spring tides, starting from high water level (HWL), continuing documented and deserves more attention in future research in the region
with mid-ebb (ME), low water level (LWL), and mid-flood (MF). The supported in robust field data.
pattern was similar in both climatic seasons but with generally greater
stratification in the wet season. During flooding and high tides, the
5.2. Residual and intertidal circulation in Tumaco Bay
plume reaches its minimum vertical extent, and coastal circulation oc­
curs in a narrow strip of depths <10 m. In this part of the tidal cycle, the
Circulation, stratification, and shear stress are shown in Fig. 12 along
plume is shrunk near the coast, and its dispersion reduces. Therefore, the
the T2 transect that crosses the entrance of Tumaco Bay (Fig. 1). The
vertical stratification increases, reaching strongly stratified conditions.
transect extends from the southwest to the northeast of the bay.
In contrast, during ebbing and low tides, the plume expands and deepens
Fig. 12A-B shows velocities cross the transect, with positive values
up to 10 m more than during high tide. In this part of the cycle, the
seaward (i.e., toward the screen). As seen in Fig. 9 for the surface, the
intense coastal circulation occurred in a wide strip as large as 3 km. This
residual velocities are directed mostly seaward in the water column, and
resulted in greater plume dispersion and mixing, weakening vertical
the strongest currents take place at the sides of the bay. This is because
stratification to partially stratified conditions. This evidences a change
near the center the bay is shallower (see Fig. 9), then, circulation is
in the stratification of the water column in the river plume with the tidal
facilitated on the sides. The bay is partially stratified toward the south,
cycle.
given the influence of Mira River, and is vertically homogeneous around
Semidiurnal changes in stratification have been identified in other
the center and north of the bay (Fig. 12C-D). Shear stress is strong in the
ROFI systems, especially at the Rhine River (Horner-Devine et al., 2015;
northern side because of the faster flow speeds and shallower depths
Rijnsburger et al., 2016; Simpson, 1997; Souza et al., 2008). This phe­
than in the southern side (Fig. 12E-F).
nomenon has been called cross-shore tidal straining (de Boer et al.,
Analyzing the circulation and salinity structure throughout the tidal
2008) and, more recently, minor-axis tidal straining (Flores et al., 2020).
cycle (Fig. 13), it was observed that the tide strongly dominated both
Nevertheless, although there are variations in stratification with the
(Fig. 13C,D), currents were directed outside the bay throughout all bay
tidal cycle in both cases (the Rhine and Mira), the causes appear to be
entrance, with higher velocities at the sides, as for the residual circu­
different. On the Rhine, the Coriolis acceleration generates an elliptical
lation pattern. In this part of the cycle, the tide flushes the waters of Mira
circulation of water in the tidal cycle. This circulation has opposing
River out of the bay; thus, at low water (Fig. 13E,F), salinities in the bay
directions in the water column between the surface and bottom as a
are the greatest in the cycle, confining the baroclinic gradient to the
product of the partial stratification by fluvial discharge. This induces a
southern extreme of the bay, while the center and north have oceanic
strong cross-shore shear in the water column that causes stratification
conditions. During flooding (Fig. 13G,H), the currents reverse the

12
Ó. Álvarez-Silva et al. Journal of Marine Systems 236 (2022) 103804

Fig. 13. Velocities perpendicular to entrance of Tumaco Bay, along transect T2 (Fig. 1), during a cycle of spring tides. Positive velocities are directed outside the bay,
i.e., toward the screen. All panels show isohalines every 2 g/kg. Dashed line shows plume outline.

13
Ó. Álvarez-Silva et al. Journal of Marine Systems 236 (2022) 103804

2.05 With river 2.05

5 10

5 10
Without river

30

30
50

50
2 2
5 5
5 5
5 5
1.95 1.95
Latitude (°)

70

70
5 5

5
10

10
30

30
5

5
10

10
1.9 5 1.9 5
50

50
5

5
5 5
5 5
5

5
1.85 30 5 5
1.85 30 5 5
10 5 10 5
5 5 5 5
5 5 5 5 5 5
10

10
5

5
5 5
10

10
5 5
5 5
5

5
5

5
1.8 5 1.8 5
5

5
10

10
5

5
5

5
5

5
5

5
1.75 1.75
-78.85 -78.8 -78.75 -78.7 -78.65 -78.6 -78.55 -78.85 -78.8 -78.75 -78.7 -78.65 -78.6 -78.55
Longitude (°) Longitude (°)

Fig. 14. Residual circulation patterns in Tumaco Bay during wet season (left) and dry season (right), considering the discharge of Mira River (blue arrows) and
without considering this discharge (red arrows) in the hydrodynamic model. (For interpretation of the references to colour in this figure legend, the reader is referred
to the web version of this article.)

direction all along the cross-section, again with greater magnitudes to­ filed data.
ward sideward and lesser toward the center. The Mira River plume also The horizontal extent of the Mira River plume takes place parallel to
enters with the tide, so that at high water (Fig. 13A,B) brackish waters the coast symmetrically in north and south directions (again due to the
have their greatest extent, covering the entire width of the bay in the wet weak effect of the Coriolis acceleration), without extending much across
season. the coast. The greatest variability of this extent was on a seasonal scale,
The weak density gradients within Tumaco Bay do not appear to and there was no evidence of substantial variability in the interannual
substantially effect the circulation inside the bay, as shown in Fig. 14, scenarios. Although the plume reaches Tumaco Bay, it is already very
where the circulation patterns are compared by including and excluding mixed and does not significantly affect the circulation. However, the
the Mira River discharge in the simulations. There are evident differ­ presence of the river generates important currents toward Tumaco Bay
ences in circulation patterns outside the bay, where there is a marked that bring the sediment that feeds the beaches and bar, protecting the
change in the direction of the currents. However, inside the bay, changes municipal seat from marine hazards.
between the two cases were negligible for the two climatic seasons.
Declaration of Competing Interest
6. Conclusions
The authors declare the following financial interests/personal re­
This study analyzed the marine dynamics in the Colombian South lationships which may be considered as potential competing interests:
Pacific coast using 3D hydrodynamic modelling. The model results Luis Otero reports financial support was provided by Direccion
indicate that the tide is the most crucial forcer in Tumaco Bay. The General Marítima de Colombia. Luis Otero reports a relationship with
circulation patterns were dominated by the ebb tide, favoring renewal of Direccion General Marítima de Colombia that includes: consulting or
the bay's waters and probably preventing the siltation of sediments and advisory. This research was developed in the framework of a research
pollutants. Circulation patterns in the bay changed very little over time, project where all authors participated as researchers. The research was
both seasonally and interannually, because the Mira River flows pro­ funded by the General Maritime Directorate of Colombia (DIMAR).
duced weak density gradients that did not appreciably influence the
hydrodynamics. Vertical circulation patterns were very homogeneous, Data availability
and depth determines the zones with the fastest speeds, with a greater
magnitude in deeper zones. This reveals an essential effect of the bottom No original data was measured for this research. All used data is from
on the circulation, classifying the bay as hydrodynamically shallow. secondary sources and can be requested from the owners, who are
On the other hand, near the Mira River mouth (south of Tumaco adequately specified in the paper.
Bay), the vertical density gradients were greater and varied strongly in
the tidal cycle near and across the coastline. Therefore, stratification was Acknowledgments
greater at flooding and high tides and weaker at ebbing and low tides.
This diurnal stratification variability has been reported in river plumes This research was funded by the Nariño Department Government,
at mid-latitudes, and its cause is strongly linked to elliptical water cir­ and the General Maritime Directorate of Colombia (DIMAR). Project
culations at the tidal scale due to the Coriolis acceleration. However, this BPIN 2018000030184. The support of the Research Vice-rectorate of
mechanism cannot cause the stratification variations in the Mira River Universidad del Norte (DIDI) is acknowledged.
plume because it is a tropical system in which the Coriolis acceleration is
negligible. Given the above, this phenomenon and its potential causes
should be studied more thoroughly in future research supported on more

14
Ó. Álvarez-Silva et al. Journal of Marine Systems 236 (2022) 103804

References Leonardi, N., Plater, A.J., 2017. Residual flow patterns and morphological changes along
a macro- and meso-tidal coastline. Adv. Water Resour. 109, 290–301. https://doi.
org/10.1016/j.advwatres.2017.09.013.
Alosairi, Y., Pokavanich, T., Alsulaiman, N., 2018. Three-dimensional hydrodynamic
Lesser, G.R., Roelvink, J.A., van Kester, J.A.T.M., Stelling, G.S., 2004. Development and
modelling study of reverse estuarine circulation: Kuwait Bay. Mar. Pollut. Bull. 127,
validation of a three-dimensional morphological model. Coast. Eng. 51, 883–915.
82–96. https://doi.org/10.1016/j.marpolbul.2017.11.049.
https://doi.org/10.1016/j.coastaleng.2004.07.014.
Alvarez, L.G., Jones, S.E., 2002. Factors influencing suspended sediment flux in the upper
LOICZ, 2005. Land-Ocean interactions in the coastal zone-science plan and
gulf of California. Stuar. Coast. Shelf Sci. 54, 747–759. https://doi.org/10.1006/
implementation strategy. In: IGBP Report 51 / IHDP Report.
ecss.2001.0873.
Murphy, P.L., Valle-Levinson, A., 2008. Tidal and residual circulation in the St. Andrew
de Boer, G.J., Pietrzak, J.D., Winterwerp, J.C., 2008. Using the potential energy anomaly
bay system, Florida. Cont. Shelf Res. 28, 2678–2688. https://doi.org/10.1016/j.
equation to investigate tidal straining and advection of stratification in a region of
csr.2008.09.003.
freshwater influence. OceanModel. 22, 1–11. https://doi.org/10.1016/j.
Nidzieko, N.J., 2010. Tidal asymmetry in estuaries with mixed semidiurnal/diurnal tides.
ocemod.2007.12.003.
J. Geophys. Res. Ocean. 115, 1–13. https://doi.org/10.1029/2009JC005864.
Cai, W., McPhaden, M.J., Grimm, A.M., Rodrigues, R.R., Taschetto, A.S., Garreaud, R.D.,
Pham, V.S., Hwang, J.H., Ku, H., 2016. Optimizing dynamic downscaling in one-way
Dewitte, B., Poveda, G., Ham, Y.G., Santoso, A., Ng, B., Anderson, W., Wang, G.,
nesting using a regional ocean model. OceanModel. 106, 104–120. https://doi.org/
Geng, T., Jo, H.S., Marengo, J.A., Alves, L.M., Osman, M., Li, S., Wu, L.,
10.1016/j.ocemod.2016.09.009.
Karamperidou, C., Takahashi, K., Vera, C., 2020. Climate impacts of the El
Poveda, G., Jaramillo, A., Gil, M.M., Quiceno, N., Mantilla, R.I., 2001. Seasonality in
Niño–Southern Oscillation on South America. Nat. Rev. Earth Environ. 1, 215–231.
ENSO-related precipitation, river discharges, soil moisture, and vegetation index in
https://doi.org/10.1038/s43017-020-0040-3.
Colombia. Water Resour. Res. 37, 2169–2178. https://doi.org/10.1029/
Chao, S.-Y., 1990. Tidal modulation of estuarine plumes. J. Phys. Oceanogr. 20,
2000WR900395.
1115–1123. https://doi.org/10.1175/1520-0485(1990)020<1115:TMOEP>2.0.CO;
Ralston, D.K., Geyer, W.R., Lerczak, J.A., Scully, M., 2010. Turbulent mixing in a
2.
strongly forced salt wedge estuary. J. Geophys. Res. Ocean. 115, 1–19. https://doi.
DANE, 2018. https://www.dane.gov.co/files/censo2018/informacion-tecnica/CNPV
org/10.1029/2009JC006061.
-2018-VIHOPE-v2.xls.
Restrepo, J., Lopez, S., 2007. Morphodynamics of the Pacific and Caribbean deltas of
Deltares, 2021. Delft3D-FLOW, User Manual.
Colombia, South America. J. S. Am. Earth Sci. 25, 1–21. https://doi.org/10.1016/j.
Devis-Morales, A., 2003a. Analysis of the oceanographic and meteorological conditions
jsames.2007.09.002.
of the Tumaco Bay and its relationship with global scale events. CCCP Sci. Bull. 9,
Restrepo, L., JC, 2006. Contribution of flows from the Baudó, San Juan, Patía and Mira
1–21.
rivers to the Colombian Pacific basin. CCCP Sci. Bull. 17–32.
Devis-Morales, A., 2003b. Evolution of the El Niño event 2002-2003 and effects on the
Rijnsburger, S., van der Hout, C.M., van Tongeren, O., de Boer, G.J., van Prooijen, B.C.,
Colombian Pacific Basin and the bay of Tumaco. CCCP Sci. Bull. 12–32.
Borst, W.G., Pietrzak, J.D., 2016. Simultaneous measurements of tidal straining and
Dyer, K.R., 1997. Estuaries, a Physical Introduction. Wiley, Chicester - UK.
advection at two parallel transects far downstream in the Rhine ROFI. Ocean Dyn.
Escobar, C.A., Velásquez, L., Posada, F., 2015. Marine currents in the Gulf of Urabá,
66, 719–736. https://doi.org/10.1007/s10236-016-0947-x.
Colombian Caribbean Sea. J. Coast. Res. 316, 1363–1374. https://doi.org/10.2112/
Saha, S., Moorthi, S., Wu, X., Wang, J., Nadiga, S., Tripp, P., Behringer, D., Hou, Y.T.,
jcoastres-d-14-00186.1.
Chuang, H.Y., Iredell, M., Ek, M., Meng, J., Yang, R., Mendez, M.P., Van Den
Flores, R.P., Rijnsburger, S., Horner-Devine, A.R., Kumar, N., Souza, A.J., Pietrzak, J.D.,
Dool, H., Zhang, Q., Wang, W., Chen, M., Becker, E., 2014. The NCEP climate
2020. The formation of turbidity maximum zones by minor axis tidal straining in
forecast system version 2. J. Clim. 27, 2185–2208. https://doi.org/10.1175/JCLI-D-
regions of freshwater influence. J. Phys. Oceanogr. 50, 1265–1287. https://doi.org/
12-00823.1.
10.1175/JPO-D-18-0264.1.
Sigaúque, P.J., Schettini, C.A.F., Valentim, S.S., Siegle, E., 2021. The role of tides, river
Geyer, W.R., MacCready, P., 2014. The estuarine circulation. Annu. Rev. Fluid Mech. 46,
discharge and wind on the residual circulation of Maputo Bay. Reg. Stud. Mar. Sci.
175–197. https://doi.org/10.1146/annurev-fluid-010313-141302.
41, 101604 https://doi.org/10.1016/j.rsma.2020.101604.
Geyer, W.R., Scully, M.E., Ralston, D.K., 2008. Quantifying vertical mixing in estuaries.
Simpson, J.H., 1997. Physical processes in the ROFI regime.Pdf. J. Mar. Syst. 12, 3–15.
Environ. Fluid Mech. 8, 495–509. https://doi.org/10.1007/s10652-008-9107-2.
Simpson, J.H., Souza, A.J., 1995. Semidiurnal switching of stratification in the region of
Guo, L., Brand, M., Sanders, B.F., Foufoula-Georgiou, E., Stein, E.D., 2018. Tidal
freshwater influence of the Rhine. J. Geophys. Res. 100, 7037–7044. https://doi.
asymmetry and residual sediment transport in a short tidal basin under sea level rise.
org/10.1029/95JC00067.
Adv. Water Resour. 121, 1–8. https://doi.org/10.1016/j.advwatres.2018.07.012.
Souza, A.J., Fisher, N.R., Simpson, J.H., Howarth, M.J., 2008. Effects of tidal straining on
Guo, X., Valle-Levinson, A., 2007. Tidal effects on estuarine circulation and outflow
the semidiurnal cycle of dissipation in the Rhine region of freshwater influence:
plume in the Chesapeake Bay. Cont. Shelf Res. 27, 20–42. https://doi.org/10.1016/j.
comparison of model and measurements. J. Geophys. Res. Ocean. 113, 1–10. https://
csr.2006.08.009.
doi.org/10.1029/2006JC004002.
Hodges, B.R., 2013. Hydrodynamical Modelling. In: Reference Module in Earth Systems
Spicer, P., Cole, K.L., Huguenard, K., MacDonald, D.G., Whitney, M.M., 2021. The effect
and Environmental Sciences. Elsevier Inc. https://doi.org/10.1016/B978-0-12-
of bottom – generated tidal mixing on tidally Pulsed River plumes. J. Phys.
409548-9.09123-5.
Oceanogr. 2223–2241. https://doi.org/10.1175/jpo-d-20-0228.1.
Horner-Devine, A.R., Jay, D.A., Orton, P.M., Spahn, E.Y., 2009. A conceptual model of
Stacey, M.T., Rippeth, T.P., Nash, J.D., 2012. Turbulence and Stratification in Estuaries
the strongly tidal Columbia River plume. J. Mar. Syst. 78, 460–475. https://doi.org/
and Coastal Seas, Treatise on Estuarine and Coastal Science. Elsevier Inc. https://doi.
10.1016/j.jmarsys.2008.11.025.
org/10.1016/B978-0-12-374711-2.00204-7
Horner-Devine, A.R., Hetland, R.D., MacDonald, D.G., 2015. Mixing and transport in
Valle-Levinson, A., 2008. Density-driven exchange flow in terms of the kelvin and Ekman
coastal river plumes. Annu. Rev. Fluid Mech. 47, 569–594. https://doi.org/10.1146/
numbers. J. Geophys. Res. Ocean. 113, 1–10. https://doi.org/10.1029/
annurev-fluid-010313-141408.
2007JC004144.
Hoyos, N., Escobar, J., Restrepo, J.C., Arango, A.M., Ortiz, J.C., 2013. Impact of the
Vannitsem, S., Chomé, F., 2005. One-way nested regional climate simulations and
2010-2011 La Niña phenomenon in Colombia, South America: the human toll of an
domain size. J.Clim. 18, 229–233. https://doi.org/10.1175/JCLI3252.1.
extreme weather event. Appl. Geogr. 39, 16–25. https://doi.org/10.1016/j.
Wolter, K., Timlin, M.S., 1998. How does 1997198 rank? events. Weather 53, 315–324.
apgeog.2012.11.018.
Yang, S.L., Zhang, J., Zhu, J., 2004. Response of suspended sediment concentration to
Hu, K., Ding, P., Wang, Z., Yang, S., 2009. A 2D/3D hydrodynamic and sediment
tidal dynamics at a site inside the mouth of an inlet: Jiaozhou Bay (China). Hydrol.
transport model for the Yangtze estuary, China. J. Mar. Syst. 77, 114–136. https://
Earth Syst. Sci. 8, 170–182. https://doi.org/10.5194/hess-8-170-2004.
doi.org/10.1016/J.JMARSYS.2008.11.014.
Zarzuelo, C., López-Ruiz, A., Díez-Minguito, M., Ortega-Sánchez, M., 2017. Tidal and
Jay, D.A., Pan, J., Orton, P.M., Horner-Devine, A.R., 2009. Asymmetry of Columbia River
subtidal hydrodynamics and energetics in a constricted estuary. Stuar. Coast. Shelf
tidal plume fronts. J. Mar. Syst. 78, 442–459. https://doi.org/10.1016/j.
Sci. 185, 55–68. https://doi.org/10.1016/j.ecss.2016.11.020.
jmarsys.2008.11.015.
Zhu, Q., Wang, Y.P., Gao, S., Zhang, J., Li, M., Yang, Y., Gao, J., 2017. Modelling
Kjerfve, B., 1979. Measurement and analysis of water current, temperature, salinity, and
morphological change in anthropogenically controlled estuaries. Anthropocene 17,
density. In: Dyer, K.R. (Ed.), Estuarine Hydrography and Sedimentation. Cambridge
70–83. https://doi.org/10.1016/j.ancene.2017.03.001.
University Press, pp. 27–53.
Kjerfve, B., 1990. Manual for Investigation of Hydrological Processes in Mangrove
Ecosystems.

15

You might also like