Nothing Special   »   [go: up one dir, main page]

Remotesensing 14 04901 v2

Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

remote sensing

Article
Improving the Modeling of Sea Surface Currents in the Persian
Gulf and the Oman Sea Using Data Assimilation of Satellite
Altimetry and Hydrographic Observations
Mahmoud Pirooznia 1 , Mehdi Raoofian Naeeni 1,2 , Alireza Atabati 1,3, * and Mohammad J. Tourian 4

1 Faculty of Geodesy and Geomatics Engineering, K. N. Toosi University of Technology, Tehran 16315-1355, Iran
2 School of Engineering, University of Newcastle, University Drive, Callaghan, NSW 2308, Australia
3 Institut für Erdmessung, Leibniz Universität Hannover, 30167 Hannover, Germany
4 Institute of Geodesy, University of Stuttgart, Geschwister-Scholl-Str. 24D, 70174 Stuttgart, Germany
* Correspondence: alireza.atabati@stud.uni-hannover.de

Abstract: Sea surface currents are often modeled using numerical models without adequately ad-
dressing the issue of model calibration at the regional scale. The aim of this study is to calibrate the
MIKE 21 numerical ocean model for the Persian Gulf and the Oman Sea to improve the sea surface
currents obtained from the model. The calibration was performed through data assimilation of the
model with altimetry and hydrographic observations using variational data assimilation, where the
weights of the objective functions were defined based on the type of observations and optimized
using metaheuristic optimization methods. According to the results, the calibration of the model
generally led the model results closer to the observations. This was reflected in an improvement of
about 0.09 m/s in the obtained sea surface currents. It also allowed for more accurate evaluations of
Citation: Pirooznia, M.; Raoofian
model parameters, such as Smagorinsky and Manning coefficients. Moreover, the root mean square
Naeeni, M.; Atabati, A.; Tourian, M.J.
error values between the satellite altimetry observations at control stations and the assimilated model
Improving the Modeling of Sea
Surface Currents in the Persian Gulf
varied between 0.058 and 0.085 m. We further showed that the kinetic energy produced by sea surface
and the Oman Sea Using Data currents could be used for generating electricity in the Oman Sea and near Jask harbor.
Assimilation of Satellite Altimetry
and Hydrographic Observations. Keywords: sea surface current; data assimilation; MIKE 21; satellite altimetry; kinetic energy; optimization
Remote Sens. 2022, 14, 4901. https://
doi.org/10.3390/rs14194901

Academic Editors: Weizeng Shao,


Alexander Babanin, Changlong Guan
1. Introduction
and Jian Sun For many years, geoscientists have been interested in the study of sea surface currents
(SSCs) and their causes. As countries exchanged goods, food, etc. through the sea, the need
Received: 17 September 2022
to gain knowledge of ocean currents was felt, and they realized that, by understanding the
Accepted: 27 September 2022
process of SSCs, they could carry out economic activities more efficiently. The significance
Published: 30 September 2022
of modeling SSCs in different water bodies, such as rivers, lakes, and open seas, is crucial
Publisher’s Note: MDPI stays neutral for economic, environmental, agricultural, and military purposes [1]. Various activities,
with regard to jurisdictional claims in such as port and coast management, marine transportation and navigation, search and
published maps and institutional affil- rescue operations, tracking pollutants and oil spills, monitoring coastal water quality,
iations. and predicting energy production from sea currents, require modeling of SSCs. Detailed
understanding of the SSCs is necessary for the accurate prediction of sedimentation in
regions close to coasts and ports. Furthermore, by examining coastal currents, it is possible
to make a correct prediction of the coastline for preparing maps of coastlines and water
Copyright: © 2022 by the authors.
boundaries. It may even be possible to prevent human casualties by accurately modeling
Licensee MDPI, Basel, Switzerland.
SSCs and providing warnings to fishing and sports organizations [2].
This article is an open access article
distributed under the terms and
This study focuses on the Persian Gulf and the Oman Sea. The Persian Gulf is the
conditions of the Creative Commons
third largest bay in the world after the Gulf of Mexico and Hudson Bay and is one of the
Attribution (CC BY) license (https:// most important strategic waterways in the world from political and economic standpoints.
creativecommons.org/licenses/by/ The Oman Sea is considered one of the deep seas in the world and connects to the Persian
4.0/).

Remote Sens. 2022, 14, 4901. https://doi.org/10.3390/rs14194901 https://www.mdpi.com/journal/remotesensing


Remote Sens. 2022, 14, 4901 2 of 22

Gulf via the Strait of Hormuz. The importance of the Strait of Hormuz is evident from the
fact that an ocean liner passes through it every 6 min [3].
The fundamental Navier–Stokes equation is the basis of SSC equations, which consist
of a set of partial differential equations and a solution based on an initial boundary-value
problem [4]. Only a few researchers have been able to achieve an analytical solution for this
problem, generally far from the real nature of SSCs in oceans [5]. With the advancement of
numerical methods in computational fluid dynamics and the increase in the computational
power of systems nowadays, numerical methods, such as finite difference, finite element,
and finite volume methods, are often employed in order to solve Navier–Stokes equations
for current modeling. As an example, the two-dimensional MIKE 21 model is a commercial
numerical model marketed by the Danish Hydraulic Institute (DHI) [6]. For single grids
(regular grids) in a spherical coordinate system, continuity and momentum equations,
temperature, salinity, and density are derived using the finite difference method, while for
flexible grids (irregular grids), the finite volume method is used for spatial discretization.
There have been a number of studies conducted in the Persian Gulf region in the area
of the numerical modeling of sea currents. Chao et al. (1992) investigated Persian Gulf
currents with a three-dimensional hydrodynamic model. They identified the combined
effects of freshwater flux and wind stress as the two essential components that drive
sea currents in this region [7]. Saberi Najafi (1997) modeled a sea current in the Persian
Gulf by solving two-dimensional hydrodynamic equations based on a finite difference
numerical method. In that research, a tidal station in the Strait of Hormuz was considered
the open boundary condition [8]. Blain (2000) used a 3D numerical model to detect tidal
currents, heat flux, wind, and density in this area. Although density was responsible
for the permanent counter-clockwise SSC in the north of the Persian Gulf, it also created
other different SSC patterns in this region [9]. Kämpf and Sadrinasab (2005) used a three-
dimensional hydrodynamic model (COHERENS model) to study the sea currents of the
Persian Gulf [10]. They found that the Persian Gulf experienced a wide range of cyclonic
currents. Sabbagh Yazdi et al. (2007) solved two-dimensional hydrodynamic equations
averaged in depth based on a finite volume method. They found that the SSC changed in
different months of the year [11]. Afshar Kaveh et al. (2017) used a finite volume method to
solve 3D hydrodynamic equations on an unstructured triangular grid in the Persian Gulf.
In all those studies, calibration of the models has been neglected, and the study area merely
confined to the Persian Gulf [12].
The application of numerical models in computing ocean currents is always subject to
some errors that can affect both the performance and the accuracy of a model. Examples
include errors in properly defining the initial and boundary conditions of a model, errors in
setting atmospheric parameters of a model (wind, air temperature, etc.), and uncertainties
in the bathymetric data. These errors are often accompanied by discretization errors in the
model equations and some other errors that are in the core of modeling [13–15]. Therefore,
all models may need to be calibrated and tuned locally to obtain accurate and reliable
results [14–16].
Data assimilation using in situ data can be used to calibrate a numerical model at
local and regional scales [13]. Calibration helps a numerical model to perform better and
tries to adjust model results versus observations via an optimization problem. This fitting
may involve improving the boundary and initial conditions of a model or developing
specific criteria for the accurate calculation of model parameters, such as Smagorinsky and
Manning coefficients.
Usually, data assimilation is performed using two techniques, sequential and varia-
tional. In this study, a variational data assimilation method is used, where an objective
function is defined, and the unknowns are determined by solving an optimization problem.
It is worth mentioning that the accuracy of the results of SSC models depends on the proper
setting of the control parameters and the accuracy of the boundary values [17]. One of the
control parameters is the bed friction coefficient, which is usually set experimentally in the
MIKE 21 model. The turbulence coefficient is also set experimentally, similar to the bed
Remote Sens. 2022, 14, 4901 3 of 22

friction coefficient. These parameters cannot be measured directly, but their tailored values
for a specific region can be estimated indirectly in a calibration process [18].
In this study, a calibration of MIKE 21 is conducted over the Persian Gulf and the
Oman Sea via data assimilation with satellite altimetry and hydrographic observations.
The calibration is performed iteratively through an optimization problem. To this end, first
the geometry and the mesh of the study area must be established using appropriate mesh
generation criteria. Then, a priori setting of initial conditions, including initial values of
SSCs and sea level height (SLH), as well as the boundary conditions of the model for both
closed and open boundaries, are applied. At the closed boundary, a zero-velocity condition
with a specified sea level based on coastal tide gauge data is used. On the other hand,
in the open boundary, the SLH data from the TM-IR01 tidal model are used. Along with
the specification of the initial and boundary conditions, some other parameters should be
introduced into MIKE 21, comprising the bed friction coefficient, eddy viscosity, wind, tide,
density, precipitation, and evaporation. With all of this information, MIKE 21 is able to
numerically solve the Navier–Stokes equation to compute the SSCs and SLH over the study
area. However, due to errors in input data, which result in errors of the boundary and
initial conditions, in conjunction with the uncertainties of some model parameters, such
as Smagorinsky and Manning coefficients or bed resistance information, the output of the
model is contaminated with some errors. To deal with these errors, a calibration of MIKE 21
via an in situ current meter, a tide gauge, and satellite altimetry data is proposed. It is based
on the idea that the outputs of the MIKE 21 model, namely SSC and SLH, become closer to
the in situ observations of these parameters as measured by the current meter, altimeter
satellites, and tide gauge. To this end, an optimization problem is defined to assimilate in
situ data with the MIKE 21 model via minimization of all possible errors in the modeling
process. This includes the errors in the force field (wind force, tidal force, and bed friction),
the initial and boundary conditions, and the model or the background, as well as the
observations used in the data assimilation, such as tide gauges, current meters, and satellite
altimetry. To achieve the assimilated model, the multiobjective function is minimized using
genetic algorithms (GAs), particle swarm optimization (PSO), teaching-learning-based
optimization (TLBO), and shuffled complex evolution (SCE). As an application and final
output of this study, by using the assimilated model, the kinetic energy caused by the SSC
is determined in order to identify areas that are prone to be used for the generation of
electricity from the SSC.
Compared to the case of single-objective calibration in which only the difference
between observations and the model is optimized, one of the innovations of this research
is the use of multiobjective calibration equipped with automatic optimization methods
that utilize all the information available in the problem [19]. Another important issue in
multiobjective calibration is the problem of defining a proper weight for different objective
functions. In this study, a simple method based on the types of observations and the
statistical methods is suggested to create the weights of the objective functions.
Another innovation of this study is the use of metaheuristic optimization algorithms
to minimize the objective function, including GA [20], PSO [21], TLBO [22], and SCE [23].
Compared to classical optimization methods, metaheuristic optimization algorithms have
numerous advantages [24,25]: Most classical optimization methods are based on gra-
dients and computations of derivatives, while metaheuristic algorithms do not require
information regarding the gradient of the objective function. This process may be highly
time-consuming for complex problem, and sometimes it leads to numerical instability. In
addition, classical optimization algorithms are highly dependent on initial values, whereas
metaheuristic optimization algorithms have stochastic and random performances and
are not sensitive to initial values. Metaheuristic algorithms are more efficient in terms of
convergence and obtaining the optimal solution. They can be used to solve complex prob-
lems, nonlinear problems, and multidimensional problems, while classical methods suffer
in dealing with these problems. Metaheuristic algorithms always find a global optimal
Remote Sens. 2022, 14, 4901 4 of 22

solution, while classical optimization methods are essentially local search methods and,
therefore, usually achieve a local optimal solution.

2. Materials and Methods


2.1. Data Description
In order to run the MIKE 21 model, the initial and boundary conditions of the model,
as well as other parameters such as bed resistance, eddy viscosity, wind, tide, density,
precipitation, and evaporation, should be prepared. To this end, the Copernicus Ocean
Circulation model was used to prepare the initial conditions of SSC and SLH. In addition,
coastal tide gauge data were used for the closed boundary conditions, and the tidal model
of the National Cartography Center (NCC) of Iran (TM-IR01) was used for the open
boundary conditions [26]. Wind data were used to determine wind force, while evaporation
and precipitation data were used to determine meteorological conditions. In addition
to the above data, the computational domain of the Persian Gulf and the Oman Sea
was defined based on Global Self-Consistent, Hierarchical, High-Resolution Geography
Database (GSHHG) data and bathymetry data from the General Bathymetric Chart of the
Oceans (GEBCO) model [27]. The assimilation of observations and validation procedures
was carried out with the help of local current meter, tide gauge, and satellite altimetry
data. In Tables 1–4, the specifications of the tide gauge data, current meter stations, satellite
altimetry, and other data used in the modeling are presented.

Table 1. Specifications of tide gauge stations.

Station Country Latitude [◦ ] Longitude [◦ ] Time Period Data Source Performance


Muscat Oman 23.633 58.567 2008–2009 UHSLC 1 modeling (CB 2 )
Masirah Oman 20.683 58.867 2008–2009 UHSLC modeling (CB)
Salalah Oman 16.933 54.007 2008–2009 UHSLC data assimilation
Karachi Pakistan 24.85 67.067 2008–2009 UHSLC modeling (CB)
Khark Iran 29.31 50.33 2008–2009 NCC 3 validation
Jask Iran 25.66 57.77 2008–2009 NCC data assimilation
Chabahar Iran 25.296 60.603 2008–2009 UHSLC modeling (CB)
Rajaei Iran 27.1 56.04 2008–2009 NCC validation
Lengeh Iran 26.55 54.88 2008–2009 NCC modeling (CB)
Khomeini Iran 30.43 49.083 2008–2009 NCC data assimilation
Bushehr Iran 28.98 50.83 2008–2009 NCC data assimilation
Bahman Iran 26.95 56.28 2008–2009 NCC modeling (CB)
Point_1 NIO 4 15 55.231 2008–2009 TM-IR01 data assimilation
Point_2 NIO 15 60.45 2008–2009 TM-IR01 modeling (OB 5 )
Point_3 NIO 15 65.45 2008–2009 TM-IR01 modeling (OB)
1 2 3
University of Hawaii Sea Level Center (Uhslc.soest.hawaii.edu). Closed boundary. National Cartography
Center of Iran. 4 North Indian Ocean. 5 Open boundary.

Table 2. Specifications of current meter stations.

Station Instrument Position Period of Data Source of Data Performance


Taheri ADCP 27.63 52.36 2008–2009 PMO 6 Data assimilation
Nayband-Gulf ADCP 27.42 52.65 2008–2009 PMO validation
Nakhl-Taghi ADCP 27.49 52.57 2008–2009 PMO Data assimilation
Kangan ADCP 27.83 52.04 2008–2009 PMO Data assimilation
6 Ports and Maritime Organization of Iran.
Remote Sens. 2022, 14, 4901 5 of 22

Table 3. Specifications of satellite altimetry data.

Mission Period Source Performance


Creation of point-wise time series for
Jason 1 2008–2009 NASA, AVISO
data assimilation and validation
Creation of point-wise time series for
ENVISAT 2008–2009 ESA
data assimilation and validation

Table 4. Specification of the data used for MIKE 21.

Data Description
GEBCO, see: https:
Bathymetry //www.gebco.net/data_and_products/gridded_bathymetry_data/
(accessed on 1 January 2022).
GSHHG, see: https://www.ngdc.noaa.gov/mgg/shorelines/
Domain
(accessed on 1 January 2022).
ERA5 hourly data, see: https://cds.climate.copernicus.eu/cdsapp#!/
Wind dataset/reanalysis-era5-pressure-levels?tab=form (accessed on
1 January 2022).
Copernicus data center, see: https://cds.climate.copernicus.eu/
Evaporation cdsapp#!/dataset/reanalysis-era5-land?tab=overview (accessed on
1 January 2022); these data were used for climatology condition
Copernicus data center, see: https://cds.climate.copernicus.eu/
Precipitation cdsapp#!/dataset/satellite-precipitation?tab=overview (accessed on
1 January 2022); these data were used for climatology condition
TM-IR01 from NCC [26]; these data were used for tidal
Tidal potential
force consideration
Copernicus data center, see https://cds.climate.copernicus.eu/
cdsapp#!/dataset/satellite-sea-level-global?tab=overview (accessed
SLA 7
on 1 January 2022); these data were used for initial condition
of modeling
Copernicus data center, see:https://cds.climate.copernicus.eu/
cdsapp#!/dataset/satellite-sea-level-global?tab=overview (accessed
SSC components
on 1 January 2022); these data were used for initial condition
of modeling
7 Sea level anomaly.

It should be noted that the chart data were considered for the reference levels of SLH
observations for both coastal tide gauges and satellites altimetry. However, the reference
level of the GEBCO model was the mean sea level (MSL), which must be transferred to
the chart data. An important point regarding the open boundary conditions in the North
Indian Ocean is that the tidal components were derived from the TM-IR01 tidal model, and
the SLH was constructed based on an empirical equation (Equation (7)). As a result, two
stations in Table 1 were considered open boundary stations.

2.2. Methodology of Calibrating MIKE Model


The model used in this study was MIKE 21 with a flexible mesh. First, geometry and
depth information must be introduced into the model. In addition to geometry and depth,
the initial and boundary conditions of the model, including SLH and SSC data at the initial
time, along with wind, precipitation, and evaporation data, should be added as inputs to
the model. Geometry and depth information are among the most important information
for ocean current modeling. A flexible meshing (without structure) was used to increase
the accuracy of the calculations and reduce the computational time. For a more efficient
calculation of the SSCs, a smaller mesh size in the islands and coastal strip (about 250 m)
was selected. As fewer changes occur in offshore areas, larger mesh sizes were chosen
Remote Sens. 2022, 14, 4901 6 of 22

(about 750 m) to speed up the performance. For the initial conditions of the SLH and SSCs,
the Copernicus Data Center, along with the MSS of the TM-IR01 model, were used (Table 4).
The closed boundary conditions were introduced as zero velocity for the SSC components
and the specified value for SLH. In fact, the SLH observations at the Chabahar, Rajaei,
Khark, Karachi, and Misirah stations were used as boundary conditions for the closed
boundary (Table 1). For open boundary conditions in the North Indian Ocean, the SLH
was determined using the TM-IR01 tidal model (Table 1). TM-IR01 is a local tide model
that is provided by the National Cartography Center (NCC) of Iran based on observations
of three altimeters, including TOPEX/POSIDON, Jason1, and Jason 2, as well as 13 coastal
tide gauge stations [26].
After creating the geometry and meshing of the model, as well as preparing the initial
and boundary conditions, the following parameters should be considered in the MIKE
21 model [28,29]:
• Selecting the equation solution as “discretization of equations with high degrees”;
• Considering density as a function of temperature and salinity;
• Variable Coriolis force in place;
• Variable wind force;
• Tidal force by applying TM-IR01 model tidal components;
• Precipitation and evaporation;
• The Smagorinsky formula was used for horizontal eddy viscosity [30], with a constant
value of 0.28 for the Smagorinsky coefficient as the suggested number for the MIKE
21 model;
• In the MIKE 21 model, one of the most valuable features is the ability to dynamically
adjust the domain of the computations. This enables one to calculate SSCs in areas
that sometimes are dry and sometimes are wet, such as tidal zones;
• Manning’s coefficient of 0.03 was applied for the bed resistance
After the above preparations, the time step for the numerical solution of the Navier–
Stokes equation should be introduced into the program. This actually depends on the type
of numerical method, as well as the dimensions and size of meshes used for discretiza-
tion [31]. The interval chosen for the time step was between 0.01 and 1800 s in accordance
with the stability criteria of the model, and the time step of running the model was 60 s.
After introducing the required parameters into the MIKE 21 model, one could run
the model to compute the SSCs and SLH. However, as it was discussed, here we tried to
calibrate the model using a variational data assimilation procedure with the aid of local
data, including coastal tide gauges (Table 1), current meters (Table 2), and satellite altimetry
data (Table 3). In this case, a weighted objective function was defined, and that function
was minimized using various optimization techniques [32]. The objective function J, based
on the errors of force field (ef ), initial conditions (ei ), boundary conditions (eb ), in situ
measurements (e), and the model (em ), was defined as follows [33]:

J( x a , F, I, B) = C f ( F − f )2 + Ci ( I − i )2 + Cb ( B − b)2 + Cy (y − H ( x m ))2 + Cm ( x a − x m )2 (1)


| {z } | {z } | {z } | {z } | {z }
ef ei eb e em

In the above equation, F is the estimated force field, f is the initial force field (including
the wind force, tides, and friction force parameters), Cf is the covariance matrix of the
force field, I is the estimated initial conditions, i is the initial conditions (including the
SLH and SSC components), Ci is the covariance matrix of the initial conditions, B is
the estimated boundary conditions, b is the boundary conditions (including SLH data),
Cb is the covariance matrix of the boundary conditions, y is the observations for data
assimilation (including SLH and SSC components), H is the interpolation operator or
observation operator, Cy is the covariance matrix of the data assimilation observations, xa is
the assimilated model, Cm is the covariance matrix of the model, and xm is the output of the
numerical model, including the SLH and SSC components derived from the MIKE model.
Remote Sens. 2022, 14, 4901 7 of 22

The error ef may arise from errors in surface wind forcing. It may also arise from sim-
plifications in equations of motion, such as hydrostatic approximation or parametrization
of turbulent mixing. The initial and boundary errors (ei and eb ) usually arise from errors in
observation. Errors in the model occur due to discretization and numerical calculations,
which have a solely mathematical nature [33].
In the next step, the covariance matrices for the force field (Cf ) [33], initial and boundary
conditions (Ci and Cb ), observations (Cy ), and model or background (Cm ) were constructed.
For the covariance matrix of the force field, including the covariance matrix of the wind
(CWind Force ), the covariance matrix of the tidal force (CTidal Potential ), and the covariance
matrix of the bed resistance (CBed Resistance ), one could write:
 
CWind Force 0 0
Cf =  0 CTidal Potential 0  (2)
0 0 CBed Resistance

Thus, we first needed to formulate the covariance matrix of the wind observations
in the study area, which were in the form of a gridded time series. In addition, for
CTidal Potential , we used the standard deviation of the tidal components, which were spec-
ified in the TM-IR01 model. The unit weight was considered for CBed Resistance . To this
end, the covariance matrix of the wind was formed using a spatio-temporal empirical
covariance function:
b∆s d∆t
CWind (∆s, ∆ t) = |ae{z } ce
| {z }
Cs Ct

in which a, b, c, and d are constant coefficients, and ∆s, ∆t are the spatial and temporal
steps, respectively [34].
In order to determine the covariance matrix of the initial conditions, which included
the SLH observations (Csea level ) and the horizontal components of the SSC (Cu and Cv ),
(the initial conditions (SLH, u, and v) were in the form of a gridded file at the initial time),
we used a spatial empirical covariance function (Cs ) to determine the covariance of each
observation [35]. Thus, one could write:
 
Csea level 0 0
Ci =  0 Cu 0  (3)
0 0 Cv

To form the covariance matrix of the boundary conditions (Cb in Equation (1)), we set
Cb = CTG , namely it was equal to the covariance matrix of the tide gauge observations. As it
was discussed, the SLH observations from tide gauge stations were used as boundary con-
ditions. To this end, for the covariance matrix of the tide gauge observations, the average of
the SLH time series (MSL) was first subtracted from the SLH time series (SLA = SLH-MSL,
which is a residual time series), and then its standard deviation was considered as the
covariance matrix of the boundary conditions (Cb ).
Finally, the covariance matrix of the observations in the data assimilation procedure,
including satellite altimetry observations (Csea level ALT ), coastal tide gauge observations
(Csea level TG ), and in situ current meter observations (Cu and Cv ), was considered as follows:
 
Csea level ALT 0 0 0
 0 Csea level TG 0 0
Cy =   (4)
 0 0 Cu 0
0 0 0 Cv

Since altimetry, tide gauge, and current observations were in the form of time series,
matrices Csea level ALT, Csea level TG , Cu , and Cv could be formed by taking the standard
deviation of their residual time series in a similar manner to explained for Cb . As it was
Remote Sens. 2022, 14, 4901 8 of 22

seen, the computation of Cu and Cv for Cy was different from the computation of Cu and Cv
for Ci .
Identifying the covariance matrix of the model or background was another challenge
of the present study. During the data assimilation process, it is very important to have a
correct estimation of this matrix. In situ observations are used to control how information
from the model space is spread out by the components of the background covariance
matrix [36]. In terms of size, when a model error is large, a greater weight is given to
the data assimilation observations. In order to form the covariance matrix of a model,
various methods have been proposed. In this study, the covariance matrix of the model
was determined using the NMC method [37]. The variance–covariance matrix of the model
was built through the comparison of 48 h forecasts of the model with 24 h forecasts of the
model at a certain time (that is, using different initial and boundary conditions at different
times for forecasting at the same specific time) [32]. Thus, one could write:

T
Cm = ( x48 − x24 )( x48 − x24 )
x48 = x True + ε48 + b48 (5)
x24 = x True + ε24 + b24

In the above relation, x48 and x24 are, respectively, the 48 h and 24 h predictions of
the model at a certain time; the upper bar indicates averaging over time or space; xtrue is
the correct value of the model (without error and bias); and ε48 and ε24 are the respective
random errors. We also assumed that there was no bias or that the bias was constant over
time (b48 = b24 ). As a result, the forecast difference could be expressed as follows [37]:

x di f f = x48 − x24 = ε48 − ε24 (6)

All of the above covariance matrices were considered as the initial values, and they
were updated in an iterative way through the calibration and optimization procedures.
In order to perform the process of calibration and data assimilation of the observations,
we used MATLAB version 9.6 toolbox from the DHI company. This toolbox is able to link
all output and input files of the MIKE 21 model to MATLAB software. In this way, by
connecting the inputs and outputs of the MIKE model, the optimization process of the
model was performed. MATLAB was used to create the objective function of Equation (1)
in an iterative process by changing the input values of the MIKE model and rerunning the
model until the model achieved acceptable accuracy [38].

3. Results
3.1. Numerical Ocean Model and Data Assimilation
In this section, the numerical results of the MIKE 21 model and data assimilation (cali-
bration) procedure using metaheuristic optimization methods are presented. Figures 1 and 2,
for a typical example, show the SLH and horizontal SSC components before and after
Remote Sens. 2022, 14, x FOR PEER REVIEW  10  of  cali-
23 
 
bration of the model on 1 January 2008.

     
(a)  (b)  (c) 
Figure 1. (a) Sea level height before data assimilation (calibration); (b) east–west component (u) of 
Figure 1. (a) Sea level height before data assimilation (calibration); (b) east–west component (u) of
the SSC before data assimilation; (c) north–south component (v) of the SSC before data assimilation. 
the SSC before data assimilation; (c) north–south component (v) of the SSC before data assimilation.
     
(a)  (b)  (c) 
Remote Sens. 2022, 14, 4901 9 of 22
Figure 1. (a) Sea level height before data assimilation (calibration); (b) east–west component (u) of 
the SSC before data assimilation; (c) north–south component (v) of the SSC before data assimilation. 

Remote Sens. 2022, 14, x FOR PEER REVIEW      10  of  23   


  (a)  (b)  (c) 
Figure 2. (a) Sea level height after data assimilation (calibration); (b) east–west component (u) of the 
Figure 2. (a) Sea level height after data assimilation (calibration); (b) east–west component (u) of the
SSC after data assimilation; (c) north–south component (v) of the SSC after data assimilation. 
SSC after data assimilation; (c) north–south component (v) of the SSC after data assimilation.

Three goals were pursued here: (i) choosing the appropriate optimization method for
calibration; (ii) selecting the correct weights; and (iii) determining the effect of the defined
objective function on the results. For the first goal, the performance of each optimization
method was evaluated following the assimilation of coastal tide gauge and current meter
data. Comparisons were made between the assimilated model, the coastal tide gauges, and
the  current meter observations, both with   and without the variance covariance  matrices
(a)  of the objective functions (Equation (1)). In Figure 3, the (c) 
(b)  mean correlation coefficient
and root mean square error (RMSE) are depicted for each optimization method in all the
Figure 1. (a) Sea level height before data assimilation (calibration); (b) east–west component (u) of 
tide gauge stations (except control stations) and current meter stations (except control
the SSC before data assimilation; (c) north–south component (v) of the SSC before data assimilation. 
stations), respectively, with and without considering the variance–covariance matrices. The
results of this study indicate that, in general, the GA and TLBO methods provided better
performances (more correlation and fewer errors compared to local data) than the PSO
and SCE methods. In addition, considering the variance–covariance matrices increased
the correlation between the assimilated model and local data. Particularly, the correlation
coefficients between the assimilated model and coastal tide gauge observations, respectively, 
in the GA, PSO, TLBO, and SCE methods(a)  without considering the variance–covariance
matrices were equal to 0.66, 0.58, 0.65, and 0.55; by considering them, they were equal to
0.82, 0.75, 0.87, and 0.69. The correlation coefficients between the assimilated model   and
   
(a)  the current meter observations,
(b)  both without considering the variance–covariance
(c)  matrices
and considering them in the GA, PSO, TLBO, and SCE methods, were equal to 0.71, 0.68,
Figure 2. (a) Sea level height after data assimilation (calibration); (b) east–west component (u) of the 
0.72, and 0.64 and 0.85, 0.79, 0.83, and 0.71, respectively.
SSC after data assimilation; (c) north–south component (v) of the SSC after data assimilation. 

 
(a) 

Figure 3. Cont.
Remote Sens. 2022, 14, 4901 10 of 22
Remote Sens. 2022, 14, x FOR PEER REVIEW  11  of  23 
 

Remote Sens. 2022, 14, x FOR PEER REVIEW  11  of  23 


 

 
(b) 
Figure 3. (a) The performance of each optimization method at tide gauge stations with and without 
Figure 3. (a) The performance of each optimization method at tide gauge stations with and without
considering covariance matrices. (b) The performance of each optimization method at current meter 
considering covariance matrices. (b) The performance of each optimization method at current meter
stations with and without considering covariance matrices. 
stations with and without considering covariance matrices.

Further investigation of the performances of the mentioned optimization methods


was provided by comparing the assimilated model with those of the control stations at
coastal tide gauge (Rajaei and Khark) and current meter stations (Nayband station). In the
GA, PSO, TLBO, and SCE methods, the RMSE values between the assimilated model and
the two control stations (Rajaei and Khark) were equal to 0.082, 0.090, 0.071, and 0.11 m
and 0.091, 0.10, 0.086, and 0.13 m, respectively. For the east–west component of the SSC, the
RMSE values were equal to 0.084, 0.088, 0.069, and 0.097 m/s, while for the north–south  
component, they were equal to 0.081, 0.085,(b)  0.073, and 0.10 m/s. As a typical example,
Figure 4 illustrates the estimation of the SLH at the Rajaei tide gauge station using each of
Figure 3. (a) The performance of each optimization method at tide gauge stations with and without 
the optimization methods. According to the obtained results, the TLBO method performed
considering covariance matrices. (b) The performance of each optimization method at current meter 
better than the other methods.
stations with and without considering covariance matrices. 

 
Figure 4. SLH at Rajaei tide gauge station using different optimization methods. 

       
(a)  Figure (b) gauge station using different optimization(c) 
Figure 4. SLH at Rajaei tide gauge station using different optimization methods. 
4. SLH at Rajaei tide methods.
Figure 5. Correlations between the observations of the SLH at the Rajaei tide gauge station and the 
For the second goal, the effect of the variance–covariance matrices (Equation (1))
model output in three cases: (a) before data assimilation, (b) after data assimilation without consid‐
on data assimilation was investigated. To this end, the process of data assimilation was
ering the variance–covariance matrices, and (c) considering the variance–covariance matrices . 
performed once considering the variance–covariance matrices and then without considering
the variance–covariance matrices (unit weight). When the variance–covariance matrices

 
stations with and without considering covariance matrices. 

Remote Sens. 2022, 14, 4901 11 of 22

were used, the correlation between the observations and the assimilated model increased,
and the assimilated model became closer to the observed data. Figure 5, for a typical
example, shows the correlations between the SLH observations at the Rajaei tide gauge
station and the model in three cases: (a) before data assimilation, (b) after data assimilation
without considering the variance–covariance matrices, and (c) considering the variance–
covariance matrices. Figure 6, for a typical example, shows the correlations between the
horizontal SSC components (u, v) at the Taheri current meter station   and the model in three
cases similar to those in Figure 5.
Figure 4. SLH at Rajaei tide gauge station using different optimization methods. 

     
(a)  (b)  (c) 
Figure 5. Correlations between the observations of the SLH at the Rajaei tide gauge station and the 
Figure 5. Correlations between the observations of the SLH at the Rajaei tide gauge station and
model output in three cases: (a) before data assimilation, (b) after data assimilation without consid‐
Remote Sens. 2022, 14, x FOR PEER REVIEW 
the model output in three cases: (a) before data assimilation, (b) after data assimilation12  of  23 
without
  ering the variance–covariance matrices, and (c) considering the variance–covariance matrices . 
considering the variance–covariance matrices, and (c) considering the variance–covariance matrices.

     
(a)  (b)  (c) 

     
(d)  (e)  (f) 
Figure 6. The correlations between the east–west component (u) of the SSC at Taheri current meter 
Figure 6. The correlations between the east–west component (u) of the SSC at Taheri current me-
station and the model in three cases: (a) before data assimilation (calibration), (b) after data assimi‐
ter station and the model in three cases: (a) before data assimilation (calibration), (b) after data
lation without considering the variance–covariance matrices, and (c) after considering the variance–
assimilation without considering the variance–covariance matrices, and (c) after considering the
covariance matrices. The correlations between the observations of the north–south component (v) 
variance–covariance matrices. The correlations between the observations of the north–south com-
of the SSC at the Taheri current meter station and the model in three cases: (d) before data assimila‐
tion, (e) after data assimilation without considering variance–covariance matrixes, and (f) consider‐
ponent (v) of the SSC at the Taheri current meter station and the model in three cases: (d) before
ing variance–covariance matrixes. 
data assimilation, (e) after data assimilation without considering variance–covariance matrixes, and
(f) considering variance–covariance matrixes.
Table 5. RMSE values between the assimilated (calibrated) model and tide gauge observations at 
Rajaei and Khark control stations and current meter observation at Nayband control station. 

  Different State of Objective Function
Cf  Cf ; Cf  Cf ; Cf  Cf ; Cf  Cf ; Cf  Cf ; C f  I;
Remote Sens. 2022, 14, 4901 12 of 22

For the third goal, to further investigate the objective function (Equation (1)) and
the performances of the variance–covariance matrices, each variance–covariance matrix
was applied separately during the data assimilation process. Table 5 shows the root mean
square error (RMSE) values between the assimilated model (calibrated model) and the
observations at control stations, including the SLH at the Rajai and Khark stations and
the SSC at the Nayband station. According to the obtained results, the errors between
the observations and assimilated models were reduced when all the variance–covariance
matrices were used. As shown in this issue, applying each of the variance–covariance
matrices improved the data assimilation process.

Table 5. RMSE values between the assimilated (calibrated) model and tide gauge observations at
Rajaei and Khark control stations and current meter observation at Nayband control station.

Different State of Objective Function


Cf = Cf ; Cf = Cf ; Cf = Cf ; Cf = Cf ; Cf = Cf ; C f = I;
Ci = Ci ; Ci = Ci ; Ci = Ci ; Ci = Ci ; Ci = I; Ci = I;
Observations Cb = Cb ; Cb = Cb ; Cb = Cb ; Cb = I; Cb = I; Cb = I;
Cy = Cy ; Cy = Cy ; Cy = I; Cy = I; Cy = I; Cy = I;
Cm = Cm ; Cm = I; Cm = I; Cm = I; Cm = I; Cm = I;
SLH at Rajaei station (meter) 0.071 0.079 0.087 0.098 0.118 0.124
SLH at Khark station (meter) 0.086 0.091 0.097 0.109 0.127 0.132
The east–west component of SSC at
0.069 0.076 0.088 0.091 0.105 0.114
Nayband station (U) (meters per second)
The north–south component of SSC at
0.073 0.087 0.094 0.103 0.110 0.119
Nayband station (V) (meters per second)

3.2. Validation and Control of Calibrated Model


As part of the validation process, we used control stations that were not involved
in the modeling and data assimilation process. Figure 7 shows the validation stations
and the names of the areas that are mentioned in this section. In order to control the
calibrated model, the Nayband current meter station, Rajai tide gauge, and Khark tide
gauge were used, along with satellite altimetry observations (167 points). It should be
mentioned that satellite altimetry observations were used for both data assimilation and
control of the results (validation). As shown in Figure 8, the study area included control
points of two missions of satellite altimetry passes, as well as coastal tide gauges and in
situ current meters. The RMSE values between the assimilated model and the observations
of control stations at the Rajai and Khark coastal tide gauges were, respectively, 0.071 m
and 0.086 m; at the Nayband current meter station, the values were 0.069 m/s for the
east–west component and 0.073 m/s for the south–north component. Comparing the RMSE
values before and after assimilation showed that the assimilation led to an improvement of
about 0.09 m/s in the SSC of the Nayband control station and about 0.25 and 0.37 m in the
SLH of the Rajai and Khark control stations, respectively. In addition, the RMSE values
between the assimilated model and the satellite altimetry observations at control stations
varied between 0.058 and 0.085 m. Moreover, the RMSE values between satellite altimetry
observations and the assimilated model at the control points varied between 0.058 and
0.085 m. A typical example is shown in Figure 9, which illustrates SLH observations at the
Rajaei control station and current meter observations at the Nayband control station, as
well as for the model before and after data assimilation.
In addition, the effect of the Manning bed resistance coefficient, the Smagorinsky
turbulence coefficient, and the amplitude and phase of the tidal components on SLH were
studied in order to examine the effect of data assimilation in the calibration of the SSC
model. There was a difference between the model and the observations when the Manning
and Smagorinsky coefficients were applied incorrectly, as well as the amplitude and phase
of tidal components. Figure 10 shows the performances of Manning coefficients of 0.015,
Remote Sens. 2022, 14, 4901 13 of 22

0.021, and 0.023 at the Rajaei control station and their effects on SLH. The optimal coefficient
at this station was 0.021. Figure 11 illustrates the optimal Manning coefficient in the study
area. Figure 12 shows the performances of Smagorinsky coefficients of 0.13, 0.15, and 0.17
at the Rajaei control station and their effects on SLH, and the optimal coefficient for this
Remote Sens. 2022, 14, x FOR PEER REVIEW  14  of  23 
station was equal to 0.15. Figure 13 shows the optimal Smagorinsky coefficient in the14 
Remote Sens. 2022, 14, x FOR PEER REVIEW 
Remote Sens. 2022, 14, x FOR PEER REVIEW 
  study
14  of 
of 23 
23 
  
area, which had smooth variations in the Persian Gulf.

 
  
Figure 7. The validation stations, along with the map of locations mentioned in the study area. 
Figure 7. The validation stations, along with the map of locations mentioned in the study area. 
Figure 7. The validation stations, along with the map of locations mentioned in the study area. 
Figure 7. The validation stations, along with the map of locations mentioned in the study area.

 
  
Figure 8. Two missions of satellite altimetry passes, coastal tide gauges, and current meter stations 
Figure 8. Two missions of satellite altimetry passes, coastal tide gauges, and current meter stations 
Figure 8. Two missions of satellite altimetry passes, coastal tide gauges, and current meter stations 
Figure 8. Two missions of satellite altimetry passes, coastal tide gauges, and current meter stations
used in the study area. 
used in the study area. 
used in the study area. 
used in the study area.

 
      
     
(a)  (b)  (c) 
(a) 
(a)  (b) 
(b)  (c) 
(c) 
Figure  9.  SLH  and  SSC  observations  at  Rajaei  and  Nayband  control  stations  obtained  from  the 
Figure 
Figure 9.9. 
9. SLH 
SLH and 
and SSC 
SSC observations 
observations at 
at Rajaei  and 
Rajaei and Nayband 
and Nayband control 
Nayband control stations 
control stations obtained 
stations obtained from 
obtained from the 
from the
the 
Figure SLH and SSC observations at Rajaei
model before and after data assimilation. (a) SLH observations at Rajaei control station (blue), the 
model before and after data assimilation. (a) SLH observations at Rajaei control station (blue), the 
model before and after data assimilation. (a) SLH observations at Rajaei control station (blue), the 
model before data assimilation (green), and the model after data assimilation (red). (b) East–west 
model before and after data assimilation. (a) SLH observations at Rajaei control station (blue), the
model before data assimilation (green), and the model after data assimilation (red). (b) East–west 
model before data assimilation (green), and the model after data assimilation (red). (b) East–west 
components of SSC at Nayband control station (blue), the model before data assimilation (green), 
model before data assimilation (green), and the model after data assimilation (red). (b) East–west
components of SSC at Nayband control station (blue), the model before data assimilation (green), 
components of SSC at Nayband control station (blue), the model before data assimilation (green), 
and the model after data assimilation (red). (c) North–south components of SSC at Nayband control 
components of SSC at Nayband control station (blue), the model before data assimilation (green),
and the model after data assimilation (red). (c) North–south components of SSC at Nayband control 
and the model after data assimilation (red). (c) North–south components of SSC at Nayband control 
andstation  (blue), 
the (blue), 
model the data
model  before  data  assimilation  (green),  and  the  model  after  data  assimilation 
station 
station  (blue), after
the 
the model 
model  assimilation
before  (red).
before data 
data  (c) North–south
assimilation 
assimilation  (green), components
(green),  and 
and the  of SSC
the model 
model  at data 
after 
after  Nayband control
data assimilation 
assimilation 
(red). 
(red). 
station
(red).  (blue), the model before data assimilation (green), and the model after data assimilation (red).
Remote Sens. 2022, 14, x FOR PEER REVIEW 
Remote 1415  of  23 
  Sens. 2022, 14, 4901
Remote Sens. 2022, 14, x FOR PEER REVIEW  15  of
of 22
23 
 
Remote Sens. 2022, 14, x FOR PEER REVIEW  15  of  23 
 

 
   
(a)  (b) 
(a) 
(a)  (b) (b) 
Figure 10. (a) Performances of Manning coefficients (MCs) of 0.015, 0.021, and 0.023 at Rajaei control 
Figure 10. (a) Performances of Manning coefficients (MCs) of 0.015, 0.021, and 0.023 at Rajaei control 
Figure 10. (a) Performances of Manning coefficients (MCs) of 0.015, 0.021, and 0.023 at Rajaei control 
station and their effects on SLH; (b) separated portion of (a) to show more details. 
Figure 10. (a) Performances of Manning coefficients (MCs) of 0.015, 0.021, and 0.023 at Rajaei control
station and their effects on SLH; (b) separated portion of (a) to show more details. 
station and their effects on SLH; (b) separated portion of (a) to show more details. 
station and their effects on SLH; (b) separated portion of (a) to show more details.

   
 
Figure 11. Optimum Manning coefficient in the study area. 
Figure 11. Optimum Manning coefficient in the study area. 
Figure 11. Optimum Manning coefficient in the study area.
Figure 11. Optimum Manning coefficient in the study area. 

   
   
(a)    (b)   
(a)  (b) 
(a)  (b) 
Figure 12. (a) Performances of Smagorinsky coefficients (SCs) of 0.13, 0.15, and 0.17 at Rajaei control 
Figure 12. (a) Performances of Smagorinsky coefficients (SCs) of 0.13, 0.15, and 0.17 at Rajaei control 
station and their effects on SLH; (b) separated portion of (a) to illustrate more details. 
Figure 12. (a) Performances of Smagorinsky coefficients (SCs) of 0.13, 0.15, and 0.17 at Rajaei control
Figure 12. (a) Performances of Smagorinsky coefficients (SCs) of 0.13, 0.15, and 0.17 at Rajaei control 
station and their effects on SLH; (b) separated portion of (a) to illustrate more details. 
station and their effects on SLH; (b) separated portion of (a) to illustrate more details.
station and their effects on SLH; (b) separated portion of (a) to illustrate more details. 

Regarding tidal components, the TM-IR01 model was used to calculate a tidal potential
for MIKE 21. The results of the TM-IR01 model have already been validated with global

 
 
 
Remote Sens. 2022, 14, 4901 15 of 22

tide models, such as the FES tide model [26]. Here, after data assimilation, we were looking
for whether the tidal components of the TM-IR01 model were improved or not. The answer
was yes. The amplitude and phase of the tidal components of this model were improved.
To show this, with the aid of an empirical tide model (Equation (7)) [39], a comparison
between the TM-IR01 tidal model and the calibrated (assimilated) model was conducted;
to achieve this, an SLH time series of the tide gauge stations was reconstructed using tidal
components of both the TM-IR01 and assimilated models, and then the differences between
the tide gauge observations and both reconstructed SLH time series were calculated for
RMSE computation.
m
SLH (t) = MSL + ∑ Ak cos(2π f k t + ϕk ) (7)
k =1

In this equation, SLH is the sea level height observed, MSL is the mean sea level, Ak is
the amplitude of the tidal components, FK is the frequency of the tidal components, ϕk is
the phase of the tidal components, and m is the number of tidal frequencies. As an example,
Figure 14 shows the SLH observation at Khark station and the reconstruction of the time
Remote Sens. 2022, 14, x FOR PEER REVIEW 
Remote Sens. 2022, 14, x FOR PEER REVIEW  16 of 
16  of 23 
23 
   series using the TM-IR01 and assimilated models. Table 6 shows the RMSE results between
the tide gauge stations and TM-IR01 model, as well as the assimilated model.

  
Figure 13. Optimum Smagorinsky coefficient in the study area. 
Figure 13. Optimum Smagorinsky coefficient in the study area.
Figure 13. Optimum Smagorinsky coefficient in the study area. 

     
(a) 
(a)  (b) 
(b) 
Figure 14. As a typical example, (a) SLH observations at Khark station (blue) and reconstructions of 
Figure 14. As a typical example, (a) SLH observations at Khark station (blue) and reconstructions of 
Figure 14. As a typical example, (a) SLH observations at Khark station (blue) and reconstructions
SLH time series with the TM‐IR01 model (green) and the assimilated (calibrated) model (red); (b) 
SLH time series with the TM‐IR01 model (green) and the assimilated (calibrated) model (red); (b) 
ofseparated portion of part (a) to show more details. 
SLH time series with the TM-IR01 model (green) and the assimilated (calibrated) model (red);
separated portion of part (a) to show more details. 
(b) separated portion of part (a) to show more details.
     
Remote Sens. 2022, 14, 4901 16 of 22

Table 6. RMSE results between SLH of tide gauge stations and TM-IR01 model, as well as assimi-
lated model.

Station TM-IR01 Model Assimilated Model


Muscat 0.121 0.081
Masirah 0.112 0.084
Salalah 0.126 0.097
Karachi 0.109 0.078
Khark 0.132 0.094
Jask 0.113 0.081
Chabahar 0.132 0.075
Rajaei 0.129 0.088
Lengeh 0.115 0.079
Khomeini 0.127 0.082
Bushehr 0.131 0.081
Bahman 0.127 0.086

3.3. Analyzing the Results of the SSC and Its Application in Producing Renewable Energy
The relevant analysis was based on data from 1 January 2008 to 1 January 2009 since
the modeling was performed during this period. As a typical example, Figure 15 illustrates
the SSCs in the Persian Gulf and the Oman Sea after data assimilation (assimilated or
calibrated model) on 1 January 2008. Following the assimilated model showed that the
SSC pattern changed in different areas, and it became closer to in situ observations through
comparison with the model before assimilation. There were, primarily, three sources of
SSCs in the Persian Gulf: density, wind, and tides. However, the main and most important
SSC in the Persian Gulf was the cyclonic eddy that occurred when seawater entered the
Persian Gulf from the Oman Sea (Figure 15). According to this pattern of SSC, the Persian
Gulf had a higher salinity than the Oman Sea. Through the Strait of Hormuz, seawater with
typical ocean salinity entered the Persian Gulf and moved northwest, parallel to the coast
of Iran, and then changed direction to the south in the western part of the basin. During
travel along this route, the water became denser, and its salinity increased as a result of
evaporation, and finally, after traveling the southern portions, this SSC became denser and
exited the Strait of Hormuz.
As a result of the changes in various parameters, density was the dominant parameter
in the northern and central parts of the Persian Gulf. SSCs caused by the wind dominated
the northwestern part of the Persian Gulf. Figure 16 shows a typical example of the differ-
ence in density of seawater on 1 January 2008 03:00 a.m. This difference was the difference
between the density of seawater and water (density of seawater (kg/m3 )—1000 (kg/m3 )).
In addition, the range of tidal currents in the Persian Gulf is large, and their amount
exceeds one meter at all times. In the area of the Tonb-Bozorg and Tonb-Kochak Islands,
as well as the Abu Musa Island, the average SSCs in different months of the year showed
the formation of eddies in the eastern and western parts of the Strait of Hormuz. In a
similar manner to [40], water entered the Persian Gulf from the northern part of the Strait
of Hormuz and spread along the northern coast. The SSC speed in the Strait of Hormuz
was typically between 0.12 and 0.25 m/s, which is consistent with a previous study by [40].
It was common to see various eddies in the Oman Sea in the north and northwest (Pakistan
and Iran), west (along the coasts of Oman), east (along the coasts of India), and in the North
Indian Ocean. These eddies were caused by the factors mentioned above.
N k 1 N k 1 2
In the above relation,  u   and  v   are, respectively, the mean SSC components in the 
east–west and north–south directions, and E is the mean kinetic energy caused by the SSC. 
Figure 17 shows the mean kinetic energy in the study area. In the Oman Sea and near Jask 
Remote Sens. 2022, 14, 4901 port, it was possible to use the kinetic energy of the currents. In this context, it is important 
17 of 22
to note that the obtained kinetic energy values were acceptable based on the estimation of 
the error of the SSC components using in situ data (Table 5). 

Remote Sens. 2022, 14, x FOR PEER REVIEW    19  of  23 


 
     
(a)  (b)  (c)  (d) 

 
     
(e)  (f)  (g)  (h) 
Figure 15. SSCs in the Persian Gulf and Oman Sea regions after the data assimilation of observations 
Figure 15. SSCs in the Persian Gulf and Oman Sea regions after the data assimilation of observations
at 3:00 a.m. on 1 January 2008: (a) the whole area, (b) Persian Gulf, (c) Strait of Hormuz, (d) Oman 
at 3:00 a.m. on 1 January 2008: (a) the whole area, (b) Persian Gulf, (c) Strait of Hormuz, (d) Oman
Sea, (e) North Indian Ocean adjacent to Oman, (f) North Indian Ocean adjacent to India, (g) Oman 
Sea, (e) North Indian Ocean adjacent to Oman, (f) North Indian Ocean adjacent to India, (g) Oman
Sea adjacent to Pakistan, and (h) Oman Sea and North Indian Ocean. 
Sea adjacent to Pakistan, and (h) Oman Sea and North Indian Ocean.
   
(e)  (f)  (g)  (h) 
Figure 15. SSCs in the Persian Gulf and Oman Sea regions after the data assimilation of obs
at 3:00 a.m. on 1 January 2008: (a) the whole area, (b) Persian Gulf, (c) Strait of Hormuz, 
Remote Sens. 2022, 14, 4901 18 of 22
Sea, (e) North Indian Ocean adjacent to Oman, (f) North Indian Ocean adjacent to India, 
Sea adjacent to Pakistan, and (h) Oman Sea and North Indian Ocean. 

 
Figure 16. Seawater density difference on 1 January 2008 at 03:00 a.m. 
Figure 16. Seawater density difference on 1 January 2008 at 03:00 a.m.

To investigate the SLH fluctuations in the Persian Gulf and the Oman Sea, the four main
tidal components of M2, S2, K1, and O1 were considered. These four main components
were applied to the model, and two different modes were considered: (1) adding wind data
and considering density; and (2) not adding wind data and considering density. As can
be seen from the results, tides dominated the SSCs of the Persian Gulf and the Oman Sea.
According to the results, tidal currents accounted for a much higher proportion of currents
(more than 70%) than other factors causing currents (winds and density variations). As
a result of the great depth of the Oman Sea, the tidal currents were slower than those in
the Persian Gulf. Thus, the areas adjacent to the Oman Sea had lower speeds and ranges
of SSCs.
Over the last few decades, researchers around the world have focused their attention
on marine renewable energy. It is known that Iran has a long coastline and that there are
both permanent and momentary SSCs with different speeds and directions. Thus, with
the aid of kinetic energy created by SSCs, they can be used as a renewable energy source.
 
In order to accomplish this, coastal areas where kinetic energy is present in significant
Figure 17. Mean kinetic energy. 
amounts are initially identified, and then equipment with different turbines is installed
to convert this energy into electricity. It was previously possible to produce 100 watts of
  electricity
  from an SSC with a current speed of 1.2 m/s, but today, this capability can be
provided by an SSC with a speed of 60 to 80 cm per second [41]. In this study, we also
identified areas in the Persian Gulf and the Oman Sea that were likely to be utilized to
produce electrical energy from the kinetic energy of SSCs. The following equation was
  used to calculate the kinetic energy of sea currents during 2008–2009 [42]:

1 N 1 N u2 + v2
u= ∑
N k =1
uk , v = ∑
N k =1
uv , E =
2
(8)

In the above relation, u and v are, respectively, the mean SSC components in the
east–west and north–south directions, and E is the mean kinetic energy caused by the SSC.
Figure 17 shows the mean kinetic energy in the study area. In the Oman Sea and near Jask
port, it was possible to use the kinetic energy of the currents. In this context, it is important
to note that the obtained kinetic energy values were acceptable based on the estimation of
the error of the SSC components using in situ data (Table 5).
Remote Sens. 2022, 14, 4901   19 of 22

Figure 16. Seawater density difference on 1 January 2008 at 03:00 a.m. 

 
Figure 17. Mean kinetic energy. 
Figure 17. Mean kinetic energy.

4. Discussion
   
The Persian Gulf and the Oman Sea are strategically important regions where petroleum
energy is transported from this region to other parts of the world. Therefore, the study of
surface currents in this region is crucial for better understanding of the regional climate,
  the environment, and maritime transportation, as well as movements of oil and non-oil
pollution. In the oceanic domain, numerical ocean circulation models have generally been
used to solve initial boundary-value problems. In the absence of knowledge regarding
parameters and boundary values, they may suffer from calibration concerns, and their
results may not be accurate enough for local and regional applications, and they may also
require local tuning (calibration) to be accurate.
In this study, we calibrated MIKE 21 for the computation of SSCs. As a result of
this approach, we encountered three main challenges: (1) selecting the correct objective
functions, (2) appropriately tuning the weights of the objective functions, and (3) selecting
an appropriate method to optimize the objective functions. To deal with these challenges,
the following novelties and innovations were considered in this manuscript
(i): A comprehensive calibration of the MIKE 21 numerical ocean model was conducted
over the Persian Gulf and the Oman Sea via data assimilation with satellite altimetry and
hydrographic observations. The calibration dealt with errors of input data, such as errors
of boundary and initial conditions, as well as uncertainties of some model parameters, such
as Smagorinsky and Manning coefficients or bed resistance information. It also considered
the intrinsic errors arising from the numerical solution to the Navier–Stokes equation.
(ii): A novel optimization problem was defined to assimilate in situ data with the
MIKE 21 model via minimization of all possible errors in the modeling process. It was a
multiobjective optimization scheme that is solved using metaheuristic algorithms, such
as genetic algorithms (GAs), particle swarm optimization (PSO), teaching-learning-based
optimization (TLBO), and shuffled complex evolution (SCE).
(iii) To deal with one of the important issues in the multiobjective calibration regarding
the definition of proper weight functions, in this study, a simple method based on the types
of observations and the statistical methods was suggested to resolve this problem, and its
effectiveness and performance were illustrated using different examples.
(iv) We also faced the challenge of lacking low-spatial-resolution data from tide gauges
and current meters. Therefore, we decided to use satellite altimetry data in the assimilation,
which is a comprehensive and reliable source of information for the study of ocean and sea
level variations.
(v) As a result of the proposed method, the SSC and SLH estimations in the Persian
Gulf and the Oman Sea were demonstrated with a reasonable degree of accuracy, and this
approach could be easily adopted to other areas. Moreover, as an interesting and feasible
application of this study, the kinetic energy caused by the SSCs was determined using the
Remote Sens. 2022, 14, 4901 20 of 22

calibrated model, and it was analyzed to identify areas that were prone to be used for the
generation of electricity from SSCs.

5. Conclusions
In this study, the MIKE 21 two-dimensional hydrodynamic model with flexible mesh
was used for modeling SSCs in the Persian Gulf and the Oman Sea. The above model was in
operation for one year. After defining the weighted objective function, the data assimilation
was performed by minimizing the function using optimization methods, such as GA, PSO,
TLBO, and SCE. In a final step, the kinetic energy associated with SSCs was determined
using the assimilated model. Generally, data assimilation improved the estimation of the
SSC components, based on the results of the data assimilation procedure. Based on the
assimilated model, the RMSE values between the coastal tide gauges at the Rajaei and
Khark control stations and the current meter at the Nayband control station were 0.071 m
and 0.086 m, as well as 0.069 m/s for the east–west SSC component and 0.073 m/s for the
north–south SSC component, respectively. Moreover, the RMSE values between satellite
altimetry observations and the assimilated model at the control points varied between 0.058
and 0.085 m. Furthermore, the appropriate application of variance–covariance matrices
increased the accuracy of estimating the SLH and SSC components. In addition, the GA and
TLBO optimization methods provided better results than PSO and SCE. According to the
results of the optimal estimation of the bed friction coefficient, as well as the Smagorinsky
turbulence coefficient, the bed friction coefficient had a significant effect on the variation in
the SLH and SSC components in the Persian Gulf and Oman Sea. This variation may be due
to the increasing depth in the Persian Gulf and the Oman Sea. Moreover, the amplitude and
phase of the tidal components were reconstructed before and after data assimilation using
an empirical tide model, which was tested with the Khark tide gauge control station and
indicated that the amplitude and phase of the assimilated model were accurate. Further,
it was possible to utilize the kinetic energy of sea currents in the Oman Sea and near the
port of Jask by examining the kinetic energy created by the SSCs in the Persian Gulf and
Oman Sea. Moreover, it was shown that the SSC pattern changed in different areas after
data assimilation, becoming more similar to the in situ observations in the Persian Gulf
and Oman Sea. According to the SSC pattern of the Strait of Hormuz, the SSC moved from
the Oman Sea toward the Persian Gulf. Due to the difference in density between the Oman
Sea and the Persian Gulf, a cyclonic eddy formed in the Persian Gulf. In the Persian Gulf,
coastal currents were primarily caused by tides, but in the Oman Sea and the North Indian
Ocean, near Pakistan, India, and Oman, eddies were created resulting from changes in
density, wind, and bathymetry.
In the context of expanding future research in this field, other sources of data, such as
sea surface temperature and sea surface salinity, could be used in the calibration process.
In addition, the proposed data assimilation method should be tested on other numerical
ocean models. Furthermore, it is possible to check areas prone to the use of marine energy
in the study area with the aid of wave modeling.

Author Contributions: Conceptualization, M.P. and M.R.N.; methodology, M.P., M.R.N., A.A. and
M.J.T.; data curation, M.P. and A.A.; formal analysis, M.P.; funding acquisition, A.A. and M.J.T.;
software, M.P. and M.R.N.; visualization, M.P. and A.A.; writing—original draft, M.P.; writing—
review and editing, M.R.N. and M.J.T. All authors have read and agreed to the published version of
the manuscript.
Funding: This research was funded by the LUH’s open-access publishing fund (https://www.tib.
eu/de/publizieren-archivieren/open-access-finanzieren/publikationsfonds-leibniz-universitaet; ac-
cessed on 17 September 2022).
Data Availability Statement: The datasets generated and analyzed during the current study are
available from the corresponding author on reasonable request.
Remote Sens. 2022, 14, 4901 21 of 22

Acknowledgments: We would like to express our great appreciation to the reviewers for the time
they invested in editing our manuscript and their valuable comments to address their concerns for
improving the paper.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Dauji, S.; Deo, M.; Bhargava, K. Prediction of ocean currents with artificial neural networks. J. Hydraul. Eng. 2014, 21, 14–27.
[CrossRef]
2. Saha, D.; Deo, M.C.; Joseph, S.; Bhargava, K. A combined numerical and neural technique for short term prediction of ocean
currents in the Indian Ocean. Environ. Syst. Res. 2016, 5, 4. [CrossRef]
3. Sadrinasab, M. Three-dimensional flushing times of the Persian Gulf. Geophys. Res. Lett. 2004, 31, L24301. [CrossRef]
4. Shchepetkin, A.; McWilliams, J. The regional oceanic modeling system (ROMS): A split-explicit free-surface topography-following-
coordinate oceanic model. Ocean Model. 2005, 9, 347–404. [CrossRef]
5. Bryan, K. A Numerical Method for the Study of the Circulation of the World Ocean. J. Comput. Phys. 1997, 135, 154–169. [CrossRef]
6. Saputra, R.A.; Perdanawati, R.A.; Akhwadhy, R. Hydrodynamic modelling using software of MIKE 21 in the land reclamation
of Jakarta Bay. Current condition and master plan. In Proceedings of the Built Environ Sci Technol International Conference,
Surabaya, Indonesia, 18–19 September 2018. [CrossRef]
7. Chao, S.; Kao, T.W.; Al-Hajri, K.R. A numerical investigation of circulation in the Arabian Gulf. J. Geophys. Res. 1992, 97,
11219–11236. [CrossRef]
8. Saberi Najafi, S. Modeling Tide in Persian Gulf Using Dynamic Nesting. Ph.D. Thesis, University of Adelaide, Melbourne,
Australia, 1997.
9. Blain, C.A. Modeling three-dimensional thermohaline-driven circulation in the Arabian Gulf. In Estuarine and Coastal Modeling,
Proceedings of the 6th International Conference, Stanford, CA, USA, 14–16 August 2000; American Society of Civil Engineers: Reston,
VA, USA, 2000; pp. 74–92.
10. Kämpf, J.; Sadrinasab, M. The circulation of the Persian Gulf: A numerical study. Ocean Sci. Discuss. 2005, 2, 129–164. [CrossRef]
11. Sabbagh-Yazdi, S.R.; Zounemat-Kermani, M.; Kermani, A. Solution of depth averaged tidal currents in Persian Gulf on unstruc-
tured overlapping finite volumes. Int. J. Numer. Methods Fluids 2007, 55, 81–101. [CrossRef]
12. Afshar-Kaveh, N.; Ghaheri, A.; Chegini, V.; Etemad-Shahidi, A.; Nazarali, M. Evaluation of different wind fields for storm surge
modeling in the Persian Gulf. J. Coast. Res. 2017, 333, 596–606.
13. Bertino, L.; Evensen, G.; Wackernagel, H. Sequential Data Assimilation Techniques in Oceanography. Int. Stat. Rev. 2007, 71,
223–241. [CrossRef]
14. Peng, S.; Xie, L.; Pietrafesa, L.J. Correcting the errors in the initial conditions and wind stress in storm surge simulation using an
adjoint optimal technique. Ocean Model. 2007, 18, 175–193. [CrossRef]
15. Zalesny, V.; Agoshkov, V.; Shutyaev, V.; Parmuzin, E.; Zakharova, N. Numerical modeling of marine circulation with 4D variational
data assimilation. J. Mar. Sci. Eng. 2020, 8, 503. [CrossRef]
16. Sirkes, Z.; Tziperman, E.; Thacker, W. Combining data and a global primitive equation ocean general circulation model using the
adjoint method. Elsevier Oceanogr. 1996, 61, 119–145. [CrossRef]
17. Mayo, T.; Butler, T.; Dawson, C.; Hoteit, I. Data assimilation within the advanced circulation (ADCIRC) modeling framework for
the estimation of manning’s friction coefficient. Ocean Model. 2014, 76, 43–58. [CrossRef]
18. Stammer, D.; Ray, R.D.; Andersen, O.B.; Arbic, B.K.; Bosch, W.; Carrere, L.; Cheng, Y.; Chinn, D.S.; Dushaw, B.D.; Egbert, G.D.
Accuracy assessment of global barotropic ocean tide models. Rev. Geophys. 2014, 52, 243–282. [CrossRef]
19. Vrugt, J.A.; Gupta, H.V.; Bastidas, L.A.; Bouten, W.; Sorooshian, S. Effective and efficient algorithm for multiobjective optimization
of hydrologic models. Water Resour. Res. 2003, 39, 8. [CrossRef]
20. Lakshmi, V.R. Optimization of thinned dipole arrays using genetic algorithm. Int. J. Eng. Technol. 2011, 3, 658–662. [CrossRef]
21. Gao, Y.; Zhu, K. Hybrid PSO-solver algorithm for solving optimization problems. J. Comput. Appl. 2012, 31, 1648–1651. [CrossRef]
22. Rao, R.V. Optimization of multiple chiller systems using TLBO algorithm. In Teaching Learning Based Optimization Algorithm;
Springer: Cham, Germany, 2016; pp. 115–128. [CrossRef]
23. Sivamathi, C.; Vijayarani, S. Assimilation high utility itemsets using shuffled complex evolution of particle swarm optimization
(SCE-PSO) optimization algorithm, 2017. In Proceedings of the International Conference on Inventive Computing and Informatics
(ICICI), Coimbatore, India, 23–24 November 2017. [CrossRef]
24. Yang, X.S. Metaheuristic Optimization. Scholarpedia 2011, 6, 11472. [CrossRef]
25. Zhang, G.; Pan, L.; Neri, F.; Gong, M.; Leporati, A. Metaheuristic Optimization: Algorithmic Design and Applications. J. Optim.
2017, 2017, 1053145. [CrossRef]
26. Soltanpour, A.; Pirooznia, M.; Aminjafari, S.; Zareian, P. Persian Gulf and Oman sea tide modeling using satellite altimetry and
tide Gauge data (TM-IR01). Mar. Georesour. Geotechnol. 2017, 36, 677–687. [CrossRef]
27. Hall, J.K. GEBCO Centennial special issue—Charting the secret world of the ocean floor. The GEBCO project 1903–2003. Mar.
Geophys. Res. 2006, 27, 1–5. [CrossRef]
Remote Sens. 2022, 14, 4901 22 of 22

28. Wang, Q.; Peng, W.; Dong, F.; Liu, X.; Ou, N. Simulating flow of an urban river course with complex cross sections based on the
MIKE 21 FM model. Water 2020, 12, 761. [CrossRef]
29. Manson, G.K. Configuration of MIKE 21 for the simulation of nearshore storm waves currents and sediment transport: Brackley
bight Prince Edward Island. Geol. Surv. Can. 2012. [CrossRef]
30. Uddin, M.; Mallik, M. Large eddy simulation of turbulent channel flow using Smagorinsky model and effects of Smagorinsky
constants. Br. J Math. Comp. Sci. 2015, 7, 375–390. [CrossRef]
31. Russell, T.F. Stability Analysis and Switching Criteria for Adaptive Implicit Methods Based on the CFL Condition; SPE Symposium on
Reservoir Simulation: Houston, TX, USA, 1989.
32. Zalesny, V.B.; Agoshkov, V.I.; Shutyaev, V.P.; Le Dimet, F.; Ivchenko, B.O. Numerical modeling of ocean hydrodynamics with
variational assimilation of observational data. Izv.-Atmos. Ocean. Phys. 2016, 52, 431–442. [CrossRef]
33. Chua, B.; Bennett, A. An inverse ocean modeling system. Ocean Model. 2001, 3, 137–165. [CrossRef]
34. Mariella, L.; Tarantino, M. Spatial temporal conditional auto-regressive model: A new autoregressive matrix. Austrian J. Stat.
2016, 39, 223–244. [CrossRef]
35. Pirooznia, M.; Raoofian Naeeni, M. The application of least-square collocation and variance component estimation in crossover
analysis of satellite altimetry observations and altimeter calibration. J. Oper. Oceanogr. 2020, 13, 100–120. [CrossRef]
36. Lee, J.C.; Huang, X. Modelling the Background Error Covariance Matrix: Applicability Over the Maritime Continent. In Data
Assimilation for Atmospheric, Oceanic and Hydrologic Applications; Park, S.K., Xu, L., Eds.; Springer: Cham, Germany, 2021; Volume IV,
pp. 599–627. [CrossRef]
37. Wang, H.; Huang, X.; Sun, J.; Xu, D.; Zhang, M.; Fan, S.; Zhong, J. Inhomogeneous background error modeling for WRF-Var using
the NMC method. J. Appl. Meteorol. Climatol. 2014, 53, 2287–2309. [CrossRef]
38. DeChant, C. Hydrologic data assimilation: State estimation and model calibration. In Dissertations and Theses; Portland State
University ProQuest Dissertations Publishing: Portland, OR, USA, 2000; p. 172. [CrossRef]
39. Pirooznia, M.; Emadi, S.R.; Alamdari, M.N. The time series spectral analysis of satellite altimetry and coastal tide gauges and tide
modeling in the coast of Caspian Sea. Open J. Mar. Sci. 2016, 6, 258–269. [CrossRef]
40. Thoppil, P.G.; Hogan, P.J. A modeling study of circulation and eddies in the Persian Gulf. J. Phys. Oceanogr. 2010, 40, 2122–2134.
[CrossRef]
41. Dudhgaonkar, P.; Duraisamy, N.; Jalihal, P. Energy extraction from ocean currents using straight bladed cross-flow hydrokinetic
turbine. Int. J. Ocean. Clim. Syst. 2017, 8, 4–9. [CrossRef]
42. Shenoi, S.S.; Saji, P.K.; Almeida, A.M. Near-surface circulation and kinetic energy in the tropical Indian Ocean derived from
lagrangian drifters. J. Mar. Res. 1999, 57, 885–907. [CrossRef]

You might also like