Open Journal of Marine Science, 2012, 2, 131-140
http://dx.doi.org/10.4236/ojms.2012.24016 Published Online October 2012 (http://www.SciRP.org/journal/ojms)
A Process Study of the Tidal Circulation in the
Persian Gulf
Stéphane Pous1, Xavier Carton2, Pascal Lazure3
1LOCEAN/IPSL, Université Pierre et Marie Curie, Pari s, France
2LPO/IUEM, Université de Bretagne Occidentale, Brest, France
3DYNECO, IFREMER, Brest, France
Email: xcarton@univ-brest.fr
Received May 16, 2012; revised June 18, 2012; accepted July 3, 2012
ABSTRACT
A homogeneous shallow-water model with free surface is used to model the tidal circulation in the Persian Gulf. The
numerical finite-difference model includes harmonic diffusion of horizontal momentum and quadratic bottom friction, it
has a 9 km mesh size and it is forced by 7 tidal components at its southern boundary. High precision bathymetric data
are used to obtain the bo ttom topography. The numerical model is run for more than a year. The results are the follow-
ing: 1) The model accurately reproduces the tidal phase and amplitude observed at 42 tidal gauges in the region. This
accuracy is attributed to the presence of the 7 components which are able to interact nonlinearly; 2) The amphidromic
points are also well positioned by the model due to a prop er ch o ice of bathymetry. This was checked also with a simpler
geometry of the domain; 3) The tidal currents can be strong in the Straits of Hormuz and in shallow areas; thus they will
have an effect of the hydrology of the region. The residual currents are weak so that they will be negligible for the
large-scale circulation on long periods; 4) Finally, the sea-surface elevation forecast by the model is in close agreement
with in-situ measurements of pressure in the Straits, performed during the GOGP99 experiment.
Keywords: Persian Gulf; Barotropic Tide; Hydrodynamical Modeling; Comparison with Data
1. Introduction
The Persian Gulf is a Northwest to Southeast oriented
basin, with length of about 1000 km, maximum width of
350 km, average depth of 40 m and maximum depth of
120 m at the Straits of Hormuz; the straits open on the
Gulf of Oman. The surface of the Persian Gulf is about
239,000 km2 and its volume is 8780 km3. It is bounded to
the North by flat land (the delta of Iranian and Iraqi riv-
ers), to the Northeast by the Zagros mountains, and to the
Southwest by the desert of Saudi Arabia. High evapora-
tion over the Persian Gulf leads to the formation of salty
waters, called the Persian Gulf Water, which are ex-
ported into the Gulf of Oman, and which are compen-
sated by an inflow of fresher Indian Ocean Surface Water.
In the present study we will not consider the complex
thermal and haline structure of the Persian Gulf, but we
will study the tidal circulation in this gulf in a simple
framework (assuming water mass homogeneity). Further
work will address the circulation in a stratified case with
thermal and haline and wind forcing.
The Persian Gulf has an oscillation period between
21.6 and 27 hours for tidal waves (from estimates by
Defant [1]). In practice, semi-diurnal and diurnal waves
generate resonant interactions in the basin which lead to
a system of amphidromic points of Kelvin-Taylor type.
In 1988, Bogdanov [2] constructed tidal charts with the
M2, S2, K1 and O1 components, obtaining different re-
sults from those by Defant [1], especially concerning the
position of amphidromic points of semi-diurnal waves.
Several models [3-5], using independent wave compo-
nents and prescribing the free surface at the boundary of
the model domain, obtained an accurate position for the
amphidromic points, but amplitudes and phases of the
tide were not always well reproduced. Recently, the use
of satellite data assimilation, reduced the model errors
[6].
In this paper, we apply a 2D shallow water model over
the Persian Gulf and the Northwestern Indian Ocean,
forced by 7 tidal components at the southern boundary,
and we describe the resulting tidal elevations and veloci-
ties; then we compare the model results with 42 tidal
gauges recordings and with data from moorings of the
GOGP99 experiment at sea, in the Straits of Hormuz.
2. The Model
A two-dimensional shallow-water model in spherical
coordinates for a homogeneous ocean was implemented
over a large domain comprising the Northwestern Indian
C
opyright © 2012 SciRes. OJMS
S. POUS ET AL.
132
Ocean, the Persian Gulf and the Gulf of Oman (see Fig-
ure 1(b)). The grid mesh was 9 km in each direction.
Bathymetric data with a 5' × 5' resolution were provided
by Proctor for the Persian Gulf (see [7]) and elsewhere,
ETOPO2 bathymetry was used (see Figure 1(a)). Four
semidiurnal tidal components M2, S2, N2, K2 and three
diurnal components K1, O1 and P1 were forced simulta-
neously on the water height at the southern boundary of
this domain. This model was started from a state of rest
and run over 375 days (the first 10 days being excluded
from the analysis, because of dynamical adjustment of
the model from the initial conditions). This long simula-
tion allowed the nonlin ear interactions between the seven
tidal components to develop fully, and also allowed a
correct separation of these components via harmonic
analysis. This was a novel aspect of this study compared
to former studies. A test was performed on the impor-
tance of the tidal generatio n potential: its inclusion in the
model only led to very small differences from the results
without it. Therefor e the simulations pr esented below did
not include this potential. The model had quadratic bot-
tom friction proportional to gravity, to the modulus of
velocity times its vector, and inversely proportional to
the squared Strickler number (equal here to 45 m–1/3·s–1)
and to the cubic root of ocean depth (see also [8,9]) The
model also had harmonic diffusion of horizontal mo-
(a)
(b)
Figure 1. (a) Bathymetry and geography of the region of Persian Gulf and Gulf of Oman; (b) The small numbers indicate the
location of tidal gauges used in tables 1 to 3. Tide recorders M1 and M2 and ADCP mooring D1 sites are indicated by blue
circles. The large black frame indicates the boundary of the 2D model domain.
Copyright © 2012 SciRes. OJMS
S. POUS ET AL. 133
mentum with a diffusivity coefficient equal to 365 m2·s–1.
3. Tidal Circulation: Model Results and
Validation
Firstly we present the iso-amplitudes and iso-phases of
tidal harmonics M2 and K1 in the Persian Gulf, provided
by the model, in Figure 2. The two amphidromic points
for semi-diurnal tides (east of Qatar and near 28˚30'N
and 50˚E), and the single amphidromic point off Qatar
for diurnal tides are recovered. These results are in very
good agreement with the analytical results by Defant [1]
as well as with the assimilated solutions by Pontius [10].
For M2 and for K1, the model adequately reproduces
the observed large amplitudes (on the order of 60 to 80
cm for M2 and 40 cm for K1) in the appropriate areas.
For M2, these areas are the Straits of Hormuz, the region
north of Qatar, and the southeastern part of the Gulf; for
K1, these regions are the southeastern part of the Gulf
and east of Qatar. For M2, the amplitudes between Qatar
and Iran are in good agreement with observations while
earlier studies tended to underestimate the amplitudes at
this location. This good result can be attributed to the
simultaneous use of 7 tidal components which can non-
linearly interact; indeed this constraint was absent from
these earlier studies.
To validate the resu lts, 42 time-series from tide gauges
over the large domain were selected among the data at
the International Hydrographic Office; these series were
chosen under the condition that th ey be sufficiently long,
continuous and consistent with neighboring tide gauges.
Harmonic analysis was performed over the data and the
model results and the 7 tidal components, presented in
amplitude and phase, are listed in Tables 1 and 2. The
model results are in very good agreement with the ob-
servations. The differences between the model and data
are often smaller than 10 percent for the amplitude and
10 degrees for the phase. Components P1 and K2 are a
little less accurate in direction because their amplitudes
are weaker. This occurs also for a few cases of O1. It
should be noted that the accuracy is good both in the
Persian Gulf and in the Gulf of Oman. Therefore, the
model adjusts everywhere from its distant forcing.
Table 3 provides an average value of the differences
for the harmonic constants between model and data. This
Figure 2. Iso-amplitude maps in cm (a), (c) and iso-phase maps in degree (b), (d) for tidal component M2 (resp. K1) in the
Persian Gulf.
Copyright © 2012 SciRes. OJMS
S. POUS ET AL.
134
Table 1. Comparison of the harmonic amplitudes in the model (mod.) and for observation (obs.) for the four main compo-
nents M2, S2, K1 and O1. H is the amplitude in cm, g is the phase shift in degrees from Gr ee nwich.
M2 S2 K1 O1
mod. obs. mod. obs. mod. obs. mod. obs.
Place Position H/g H/g H/g H/g H/g H/g H/g H/g
1 Salalah 16˚56/54˚0030/141 31/145 11/170 12/168 33/343 36/345 18/345 18/348
2 Lakbi 18˚14/56˚3439/153 40/154 14/183 14/183 34/343 34/346 18/345 19/345
3 Sirab 20˚10/57˚4947/158 57/156 17/189 24/185 35/341 39/340 18/345 20/342
4 Rounders bay 20˚12/58˚3846/158 55/156 17/188 20/187 35/341 42/346 18/345 19/343
5 Sur 22˚34/59˚3261/162 60/165 22/193 23/196 36/341 40/351 18/345 19/342
6 Diba 23˚05/59˚0462/162 62/160 23/193 26/194 37/341 33/339 18/345 22/344
7 Muscat 23˚38/58˚3464/163 63/160 23/194 24/191 37/341 39/342 18/345 20/343
8 Saham 24˚09/56˚5469/163 68/159 25/194 26/189 37/341 40/339 18/345 22/343
9 Khor al Fakkan 25˚21/56˚2271/165 66/162 26/196 27/192 37/341 35/340 18/346 19/346
10 Ras Dillah 26˚08/56˚2874/171 72/174 27/202 27/202 34/345 31/341 18/349 19/353
11 Khor Khwair 25˚58/56˚0361/220 62/210 20/262 20/250 23/53 21/29 15/23 16/28
12 Ajman 25˚25/55˚2647/238 43/248 17/284 13/301 25/73 17/103 14/33 15/50
13 Dubai 25˚15/55˚1642/247 44/237 16/297 16/281 27/80 23/91 14/37 16/42
14 Khor Ghanada 24˚50/54˚4638/254 42/263 15/305 15/318 29/83 27/105 13/39 15/56
15 H. al Mubarr as 24˚27/53˚2232/251 28/261 14/310 13/315 33/83 43/102 14/38 23/52
16 Zarqa 24˚53/53˚0511/212 11/266 5/313 4/337 32/83 36/95 13/37 16/62
17 Ad Dawhah 25˚18/51˚3136/67 32/51 9/109 11/82 27/59 36/72 10/12 16/28
18 Ras Laffan 25˚54/51˚3544/46 38/37 13/103 11/76 16/44 25/53 5/353 12/0
19 Jabal Fuwairat 26˚03/51˚2249/46 42/44 15/106 13/88 12/33 20/54 4/333 9/0
20 Sitra 26˚10/50˚4058/73 66/58 21/137 19/120 8/342 10/347 5/275 6/264
21 Bharain 26˚22/50˚4763/57 63/54 22/119 20/102 9/326 9/10 5/262 7/268
22 Khwar Fasht 26˚20/50˚2666/50 55/62 24/113 20/114 12/293 8/343 7/241 8/256
23 Ras Tannurah 26˚39/50˚1057/38 60/42 21/100 20/98 14/277 14/294 9/229 12/238
24 Ras al Qulay’ah 26˚52/49˚5451/35 48/36 19/98 16/93 16/269 18/274 9/223 10/218
25 Berri 27˚13/49˚4343/33 44/37 17/98 16/107 18/262 17/273 10/219 14/227
26 Ras al Mishaab 28˚07/48˚3832/262 25/276 10/326 8/335 38/256 38/259 19/214 21/221
27 Mina al Ahmadi 29˚04/48˚1064/237 63/248 22/299 17/312 44/252 43/263 21/211 29/215
28 Shatt al Arab 29˚50/48˚4374/217 84/221 26/279 29/279 46/244 50/250
22/203 30/205
29 Jazh ye Khark 29˚19/50˚2030/169 36/149 10/223 13/196 36/230 39/233 18/190 26/192
30 Bushehr 28˚54/50˚4524/120 34/110 9/172 12/160 31/223 31/227 16/184 20/189
31 Lavar 28˚15/51˚1638/63 50/68 15/120 18/117 21/207 25/209 12/171 18/171
32 Asalu 27˚28/52˚3752/28 51/19 17/83 17/57 18/117 24/115 8/90 12/89
33 Jezirat Lavan 26˚48/53˚2333/350 30/335 10/34 12/10 27/91 29/94 11/58 15/55
34 Jezirat Forur 26˚15/54˚3135/273 45/259 13/316 15/295 30/78 38/86 14/43 22/42
35 Bandar Lengeh 26˚44/54˚5355/249 60/230 19/292 23/267 32/68 33/67 16/36 22/33
36 Jezirat Tunbh 26˚16/55˚1855/239 59/232 19/283 20/269 30/65 29/66 16/33 19/39
37 Henjam 26˚41/55˚5470/221 74/204 24/261 25/247 30/46 29/28 18/22 20/14
38 Bandar Abbas 27˚11/56˚1795/196 00/197 33/231 36/229 34/15 34/11 20/6 21/3
39 Pasni 25˚12/63˚3061/160 72/166 22/191 26/192 37/339 28/335 18/344 21/346
40 Karachi 24˚48/66˚5870/161 79/163 26/192 30/194 38/339 40/341 18/343 20/342
41 Porbandar 21˚38/69˚3761/161 65/180 22/192 24/211 36/337 35/347 17/342 17/350
42 Ratnagiri 16˚59/73˚18 61/170 66/175 21/203 26/209 34/335 35/340 16/343 16/349
M1 25˚14/53˚5125/267 31/262 11/316 13/296 29/81 34/90 13/39 20/43
M2 25˚36/57˚0072/165 70/161 26/196 29/190 37/343 47/335 19/346 21/347
Copyright © 2012 SciRes. OJMS
S. POUS ET AL. 135
Table 2. Comparison of the harmonic amplitudes in the model (mod.) and for observation (obs.) for the other waves N2, K2
and P1. H is the amplitude in cm, g is the phase shift in degrees from Greenwich. The last column shows F, the type of tide.
N2 K2 P1 Type F
mod. obs. mod. obs. mod. obs. mod. obs.
Place Position H/g H/g H/g H/g H/g H/g Fc. Fo.
1 Salalah 16˚56/54˚00 8/137 8/133 3/168 3/161 11/343 11/343 1.23 1.26
2 Lakbi 18˚14/56˚34 10/146 11/138 4/182 4/183 11/343 12/347 0.99 0.98
3 Sirab 20˚10/57˚49 11/149 14/138 5/188 6/185 12/342 13/340 0.82 0.73
4 Rounders bay 20˚12/58˚38 11/148 14/140 5/187 6/187 12/341 14/346 0.84 0.81
5 Sur 22˚34/59˚32 14/151 15/146 6/193 6/195 12/341 13/351 0.66 0.71
6 Diba 23˚05/59˚04 14/152 16/147 6/192 7/194 12/341 11/339 0.65 0.63
7 Muscat 23˚38/58˚34 15/152 16/141 7/192 6/184 12/341 12/340 0.63 0.69
8 Saham 24˚09/56˚54 16/152 17/139 7/193 7/188 12/342 13/339 0.59 0.65
9 Khor al Fakkan 25˚21/56˚22 17/154 13/155 7/195 7/192 12/342 12/340 0.57 0.58
10 Ras Dillah 26˚08/56˚28 17/159 18/160 8/201 7/202 11/345 11/341 0.52 0.51
11 Khor Khwair 25˚58/56˚03 14/204 11/212 6/258 5/250 7/48 7/29 0.47 0.45
12 Ajman 25˚25/55˚26 11/220 1/227 5/281 4/301 8/69 6/103 0.61 0.57
13 Dubai 25˚15/55˚16 9/229 10/217 5/293 5/265 8/77 7/77 0.71 0.64
14 Khor Ghanada 24˚50/54˚46 8/235 9/248 5/300 4/318 9/80 9/105 0.79 0.74
15 H. al Mubarr as 24˚27/53˚22 6/232 4/232 5/300 3/315 9/82 14/102 1.00
1.62
16 Zarqa 24˚53/53˚05 2/164 1/227 2/292 1/343 9/83 12/93 2.90 3.43
17 Ad Dawhah 25˚18/51˚31 9/44 9/30 3/128 3/70 8/59 10/67 0.80 1.23
18 Ras Laffan 25˚54/51˚35 10/22 10/12 5/103 3/64 5/44 7/48 0.37 0.76
19 Jabal Fuwairat 26˚03/51˚22 11/22 11/17 6/101 4/88 3/33 7/54 0.25 0.53
20 Sitra 26˚10/50˚40 12/48 13/29 8/125 5/120 2/334 3/347 0.16 0.19
21 Bharain 26˚22/50˚47 13/31 12/20 9/108 5/102 2/319 3/10 0.16 0.19
22 Khwar Fasht 26˚20/50˚26 13/25 10/36 9/102 5/114 3/285 3/343 0.21 0.21
23 Ras Tannurah 26˚39/50˚10 11/13 13/13 8/90 7/77 4/270 4/274 0.29 0.33
24 Ras al Qulay’ah 26˚52/49˚54 10/11 10/9 7/87 4/93 5/263 6/274 0.36 0.44
25 Berri 27˚13/49˚43 8/11 9/5 6/86 4/107 5/257 6/273 0.47 0.52
26 Ras al Mishaab 28˚07/48˚38 7/234 6/243 4/319 3/334 11/253 13/253 1.35 1.79
27 Mina al Ahmadi 29˚04/48˚10 13/213 12/226 9/292 5/312 13/249 14/263 0.76 0.90
28 Shatt al Arab 29˚50/48˚43 15/195 17/189 10/272 10/260 14/242 14/243 0.68 0.71
29 Jazh ye Khark 29˚19/50˚20 7/147 8/134 4/219 4/190 11/227 12/222 1.34 1.31
30 Bushehr 28˚54/50˚45 5/100 7/84 3/165 4/156 9/220 9/218 1.43 1.11
31 Lavar 28˚15/51˚16 8/41 40/44 5/111 5/117 7/203 9/210 0.62 0.64
32 Asalu 27˚28/52˚37 11/6 11/352 6/79 5/57 6/115 8/116 0.38 0.52
33 Jezirat Lavan 26˚48/53˚23 7/330 8/321 3/37 4/300 8/90 10/328 0.89 1.05
34 Jezirat Forur 26˚15/54˚31 8/257 9/253 4/313 4/295 9/76 13/86 0.94 0.99
35 Bandar Lengeh 26 ˚44/54˚53 12/232 13/212 6/289 1/265 10/64 11/62 0.65 0.66
36 Jezirat Tunbh 26˚16/55˚18 12/223 14/208 6/280 6/269 9/61 10/66 0.62 0.61
37 Henjam 26˚41/55˚54 16/206 17/178 7/259 7/247 10/42 9/28 0.51 0.50
38 Bandar Abbas 27˚11/56˚17 22/183 22/180 10/229 10/227 11/13 11/10 0.42 0.40
39 Pasni 25˚12/63˚30 14/149 17/174 6/190 7/192 12/340 9/336 0.66 0.50
40 Karachi 24˚48/66˚58 16/150 18/147 7/191 8/185 12/340 11/337 0.58 0.56
41 Porbandar 21˚38/69˚37 14/148 16/163 6/191 7/214 12/337 10/347 0.64 0.59
42 Ratnagiri 16˚59/73˚18 13/155 15/158 6/203 7/212 11/335 11/341 0.61 0.55
M1 25˚14/53˚51 5/252 6/243 3/307 4/300 9/79 11/86 1.18 1.20
M2 25˚36/57˚00 17/153 17/145 7/195 8/192 12/343 15/308 0.57 0.68
Copyright © 2012 SciRes. OJMS
S. POUS ET AL.
Copyright © 2012 SciRes. OJMS
136
Table 3. Mean value of the differences between the model and the observed harmonic amplitudes H (in cm) and phase g (in
degrees).
M2 S2 K1 O1 N2 K2 P1
Averaged difference for the: H/g H/g H/g H/g H/g H/g H/g
11 gauges of Arabian Sea 4/5 2/4 3/4 2/3 1/9 0/6 1/4
31 gauges of the Persian Gulf 5/10 2/13 3/11 4/8 3/11 1/19 1/19
42 gauges (total) 5/9 2/11 3/10 3/6 2/11 1/14 1/10
average is taken over all tidal gauges in the gulf, in the
Arabian Sea or in the whole domain. Again, the model
results agree very well with the observations. The aver-
aged differences are on the order of a few centimeters for
the amplitude and a few degrees for the phase. Very
small errors are found in the Arabian Sea. The largest
errors are found for waves K2 and P1 in the Persian Gulf;
these errors can come from both the model and the
measurements.
Another element of the model which leads to precise
results is the accurate bathymetry in the Persian Gulf (see
also [11]). Navigation charts had originally been used to
provide the bottom depth in the gulf; but these charts
were not accurate enough (the gulf was too shallow) and
they led to a misrepresentation of the amphidromic
points. This was corrected by using the bathymetry pro-
vided by Proctor and complemented with ETOPO2. The
difference between the depth of the navigation charts and
that used finally in the model was about 10 m on averag e
over the gulf.
To show how important an accurate bathymetry is for
the model, we tested the influence of bottom topography
on tides in the idealized case of a rectangular basin with
an open strait at its eastern boundary; in that case, the
tide was prescribed at this strait with the same amplitude
and phase as in the real Persian Gulf.
Indeed, Defant [1] considered the reflection of Kelvin
waves in a semi-enclosed rectangular basin, and showed
that the natural period of waves in a basin comparable to
the Persian Gulf was 22 - 23 hours. Thus, with a diurnal
or semi-diurnal forcing, resonance oscillations can ap-
pear. When the forcing is a semi-diurnal wave, two am-
phidromic points should appear.
These two points are recovered in our simple basin
experiment when the average depth is 40 m but not when
it is 30 m. This shows that a proper b athymetry is critical
for a good representation of tides.
Finally, the nature of the tide in the Gulf varies de-
pending on the location. Table 2 provide the value of
ratio F defined as F = (K1 + O1)/(M2 + S2), which char-
acterizes the type (diurnal, semi-diurnal or mixed) of tide.
For F < 0.25 the tide is semi-diurnal; for 0.25 < F < 1.5,
the tide is semi-diurnal with diurnal inequality; for 1.5 <
F < 3.0, the tide is mixed; and finally, for F > 3.0, the
tide is diurnal. Figure 3 shows the value of factor F in
the Persian Gulf. The model correctly reproduces the
three types of tides (semi-diurnal, mixed and diurnal)
observed by John [12]. Figure 4 shows the time evolu-
tion of the free surface in the 2D model at three points
with different types of tides. Point A is located near the
diurnal amphidromic point; since the diurnal component
of the free surface elevation is zero at the amphidromic
point, the semi-diurnal component dominates the tidal
signal as observed. Conversely, point B is located near
the semi-diurnal amphidromic point so that the diurnal
component dominates. Finally, point C is not close to an
amphidromic point so that the tide is mixed at this point
as shown by Figure 4.
Figure 5(a) shows the maximal velocities of the tidal
current in the Persian Gulf over the year of simulation;
these current maxima do not necessarily occur at the
same time everywhere. Tidal currents can be fast nearly
everywhere in the Gulf; a striking feature is that such fast
currents do not specifically depe nd on the type of tide. In
particular, currents faster than 1 m/s are found in the
Straits of Hormuz and around islands. Eulerian residual
currents were then computed (because the mesh size is
too large to allow the computation of Lagrangian residual
currents) at a time of maximal diurnal and semi-diurnal
currents; this period was near day 165 of the simulation.
The currents were then averaged over a 20 day period
around that date (to filter the instantaneous tidal signal).
The residual currents are weak over the whole Gulf; ve-
locities do not exceed 2 cm/s (see Figure 5(b)). An anti-
cyclonic current pattern can be seen north of Qatar with
an intensification along the Iranian coast. Faster recircu-
lations are observed along the coasts, around islands and
in the Straits of Hormuz. These results agree with those
of a finite element model by [13].
4. Validation with Complementary Data
A complementary validation was performed using data
from the GOGP99 experiment at sea (see [14,15]). This
experiment was carried out in the Straits of Hormuz and
in the Gulf of Oman in October 1999. Figure 1(b) shows
the location of the moorings M1, M2 and D1 of this ex-
periment.
D1 is a Doppler current-meter; it was moored at 120 m
depth at the exit of the Straits of Hormuz. The Doppler
S. POUS ET AL. 137
Figure 3. Map of factor F over the Persian Gulf indicating the type of tide.
Figure 4. Time evolution of the free surface deviation (in meters) in the model for three areas typical of the Persian Gulf.
Point A is located near the diurnal amphidromic point, point B is located near the westernmost semi-diurnal amphidromic
point and point C lies in the northwesternmost part of the Gulf, near the Shatt-al-Arab.
current-meter was equipped with a pressure probe and
recorder which provided a time-series. In Figure 6, this
pressure measurement is compared with the variation of
the free surface elevation simulated by the model, once
the average has been substracted from each series of data,
during the same period. The two curves are strikingly
similar both in amplitude (where the difference is a few
centimeters only) and in phase. We can assume that the
effects of the meteorological forcings at the ocean sur-
face have been approximately canceled by substracting
the 10-day average.
Furthermore, two tide gauges, M1 and M2, have been
moored at the entrance and exit of the Straits of Hormuz,
for about a month in 1999. After the experiment, a har-
monic analysis of the data was achieved and the ampli-
tude and phase of the main components of the tide were
determined at each point. These results were then com-
pared with those forecast at the same points by the nu-
merical model. It is interesting to note that these mea-
surements lie in the open sea, contrary to those which
were used above to validate the model. The agreement
between model and data is good with a difference of only
Copyright © 2012 SciRes. OJMS
S. POUS ET AL.
138
(a)
(b)
Figure 5. (a) Maximum velocities of tidal currents in the persian gulf (in m/s); (b) Residual tidal current in the persian gulf.
a few centimeters for all waves between the modeled and
measured amplitudes, and of less than 10 degrees for the
phase (see last two rows of Tables 1 and 2). Only the
phase of component P1 at mooring M2 presents a large
difference (larger than 30 degrees) between the model
and the measurements. This difference is probably due to
the fact that the model grid size (about 9 km) does not
allow an accurate sampling of the bathymetry in the mo del.
If the error remains weak in the Persian Gulf (mooring
M1 lies at 25 m depth in reality and at 32 m in the
model), this error is much larger at the exit of the Straits
where the topographic gradients are large (mooring M2
lies at 113 m depth in reality and at 203 m in the model).
5. Conclusions
We have implemented a finite-difference code of a shal-
low-water model for a homogeneous ocean. This model
was forced by 7 tidal components at its southern bound-
ary. It reproduced well the tidal elevations and currents
over the Persian Gulf and, more generally, over the larger
domain (the Northwestern Indian Ocean). The amphi-
dromic points and the type of tidal variation (diurnal,
semi-diurnal, mixed) were correctly reproduced. The
model results were validated with success using data
from 42 tidal gauges, both in the Persian Gulf and in the
Arabian Sea. This success can be attributed to an accu-
Copyright © 2012 SciRes. OJMS
S. POUS ET AL. 139
Figure 6. Measured pressure anomaly (in red, in dbars) and modeled free surface elevation (in blue, in meters) with respect
to the day of year 1999.
rate bathymetry and to the simultaneous inclusion of the
7 tidal components, which can nonlinearly interact.
The tidal currents prov ided by the model are fast in the
Straits of Hormuz (among other locations) and thus they
will efficiently contribute to mix the outflowing Persian
Gulf Water with the inflowing Indian Ocean Surface
Water. They can also be fast in shallow areas and thus
they will have an important effect on the hydrology in
the whole Persian Gulf. But the Eulerian residual tidal
currents are weak, and therefore they will be a priori
negligible in the general circulation on the long term.
Finally, the model accurately reproduced the measure-
ments of free surface elevation and of tidal components
at three moorings during the GOGP99 experiment.
This tidal model is now validated and can be included
in a 3D primitive equation model for further studies.
These studies will first address the effect of the wind on
river plumes and later the regional circulation forced by
the wind and by the atmospheric (heat and freshwater)
fluxes.
6. Acknowledgements
The authors were funded by SHOM (French Navy Hy-
drographic and Oceanographic Service), by IFREMER
(French Institute for Marine Research and Development),
by the MNHN (Museum National d’Histoire Naturelle)
and by UBO (Universite de Bretagne Occidentale). XC
thanks DGA (Defence Research Agency) for support
under REI program COMINO.
REFERENCES
[1] A. Defant, “Physical Oceanography,” Pergamon Press,
Vol. 2, 1960.
[2] K. T. Bogdanov, “Propagation des Ondes de Marees et
des Variations de Marees du Niveau du Golfe Persique,”
Marees Terrestres, Vol. 101, 1988, pp. 7004-7009
[3] L. Von Trepka, “Investigations of the Tides in the Persian
Gulf by Means of a Hydrodynamical Numerical Model,”
Proceedings of the Symposium on Mathematical Hydro-
dynamical Investigations of the Physical Processes in the
Sea, Institut fur Meereskunde der Universitat Hamburg,
Vol. 10, 1968, pp. 59-63.
[4] D. J. Evans-Roberts, “Tides in the Persian Gulf,” Con-
sulting Engineer, Vol. 43, No. 6, 1979, pp. 46-48.
[5] R. W. Lardner, M. S. Belen and H. M. Cekirge, “Finite
Difference Model for Tidal Flows in the Arabian Gulf,”
Computers & Mathematics with Applications, Vol. 8, No.
6, 1982, 425-444.
[6] L. Kantha, “Tides—A Modern Perspective,” Marine Ge-
odesy, Vol. 21, No. 4, 1998, pp. 275-297.
[7] R. Proctor, R. A. Flather and A. J. Elliott, “Modeling
Tides and Surface Drift in the Arabian Gulf: Application
to the Gulf Oil Spill,” Continental Shelf Research, Vol.
14, No. 5, 1994, pp. 531-545.
[8] S. Pous, “Dynamique Oceanique dans les Golfes Persique
et d’Oman”, Ph.D. Thesis, Universite de Bretagne Occi-
dentale, Brest, 2005, 365 p.
[9] P. Lazure and F. Dumas, “An External-Internal Mode
Coupling for a 3D Hydrodynamical Model for Applica-
tions at Regional Scale (MARS),” Advances in Water
Resources, Vol. 31, No. 2, 2007, pp. 233-250.
doi:10.1016/j.advwatres.2007.06.010
[10] P. Pontius, L. Kantha, V. Anantharaj and T. J. Bennett,
“Tidal Modeling in Marginal and Semi-Enclosed Seas,”
Proceedings of the MTS 94 Conference, Challenges and
Opportunities in the Marine Environment, Washington,
7-9 September 1994, pp. 770-776.
Copyright © 2012 SciRes. OJMS
S. POUS ET AL.
140
[11] P. C. Roos and H. M. Schuttelaars, “Influence of Topog-
raphy on Tide Propagation and Amplification in Semi-
Enclosed Basins,” Ocean Dynamics, Vol. 61, No. 1, 2011,
pp. 21-38.
[12] V. C. John, “Harmonic Tidal Current Constituents of the
Western Arabian Gulf from Moored Current Measure-
ments,” Coastal Engineering, Vol. 17, 1992, pp. 145-151.
doi:10.1016/0378-3839(92)90016-N
[13] C. A. Blain, “Barotropic Tidal and Residual Circulation
in the Arabian Gulf,” In: M. L. Spaulding and H. L. But-
ler Eds., Proceedings of the 5th International Conference
on Estuarine and Coastal Modeling, American Society of
Civil Engineers, 1998, pp. 166-180.
[14] S. Pous, X. Carton and P. Lazure, “Hydrology and Circu-
lation in the Straits of Hormuz and the Gulf of Oman;
Results from the GOGP99 Experiment. I. Straits of Hor-
muz,” Journal of Geophysical Research, Vol. 109, No.
C12037, 2004, pp. 1-15.
[15] S. Pous, X. Carton and P. Lazure, “Hydrology and Circu-
lation in the Straits of Hormuz and the Gulf of Oman;
Results from the GOGP99 Experiment. II. Gulf of Oman”,
Journal of Geophysical Research, Vol. 109, No. C12038,
2004, pp.1-26.
Copyright © 2012 SciRes. OJMS