-
Notifications
You must be signed in to change notification settings - Fork 0
/
SQGE.Tests.tex
executable file
·765 lines (708 loc) · 37.8 KB
/
SQGE.Tests.tex
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
634
635
636
637
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
681
682
683
684
685
686
687
688
689
690
691
692
693
694
695
696
697
698
699
700
701
702
703
704
705
706
707
708
709
710
711
712
713
714
715
716
717
718
719
720
721
722
723
724
725
726
727
728
729
730
731
732
733
734
735
736
737
738
739
740
741
742
743
744
745
746
747
748
749
750
751
752
753
754
755
756
757
758
759
760
761
762
763
764
765
The main goal of this section is twofold. First, we show that the FE
discretization of the streamfunction formulation of the SQGE \eqref{eqn:SQGE_Psi}
with the Argyris element produces accurate numerical approximations. To this
end, we benchmark our numerical results against those in the published
literature \cite{Vallis06, Cascon, Myers}. The second goal is to show that the
numerical results follow the theoretical error estimates in
\autoref{thm:EnergyNorm} and \autoref{thm:Errors}, i.e., we compare the observed
rates of convergence to the theoretical rates of convergence developed in
\autoref{sec:SQGEErrors}.
Although the pure streamfunction formulation of the steady SQGE
\eqref{eqn:SQGE_Psi} is our main concern, we also test our Argyris FE
discretization on two simplified settings: \begin{inparaenum}[(i)] \item the
\emph{linear Stommel} model; and \item \emph{linear Stommel-Munk} model.
\end{inparaenum} The reason for using these two additional numerical tests is
that they are standard test problems in the geophysical fluid dynamics
literature (see, e.g., Chapter 13 in Vallis \cite{Vallis06} as well as the
reports of Myers \emph{et al.} \cite{Myers}, and Cascon \emph{et al.}
\cite{Cascon}). This allows us to benchmark our numerical results against
those in the published literature. Since both the linear Stommel and the
linear Stommel-Munk models lack the nonlinearity present in the SQGE
\eqref{eqn:SQGE_Psi}, they represent good stepping stones to verifying our FE
discretization.
The first simplified model used in our numerical investigation is the
\emph{linear Stommel} model (equation (14.22) in \cite{Vallis06} and equation
(11) in \cite{Myers})
\begin{equation}
\epsilon_s \Delta \psi + \frac{\partial \psi}{\partial x} = f,
\label{eqn:Stommel}
\end{equation}
where $\epsilon_s$ is the Stommel number. The Stommel number is given by
(equation (10) in \cite{Myers})
\begin{equation*}
\epsilon_s = \frac{\gamma}{\beta L},
\end{equation*}
where $\gamma,\, \beta, \text{ and } L$ are the bottom fricition decay, the
measure of the $\beta$-plane effect, and characteristic length scale,
respectively. We note that the linear Stommel Model \eqref{eqn:Stommel} is just
the linear Stommel-Munk model \eqref{eqn:Stommel-Munk} in which the biharmonic
term is dropped (i.e. $\epsilon_m=0$). Thus it is a good starting point, since
is less complex than both the SQGE and the linear Stommel-Munk model.
The \emph{linear Stommel-Munk} model (see p. 587 in \cite{Vallis06} and problem
2 in \cite{Cascon}) is given by
\begin{equation}
\epsilon_s \Delta \psi - \epsilon_m \Delta^2 \psi + \frac{\partial \psi}{\partial x} = f,
\label{eqn:Stommel-Munk}
\end{equation}
where $\epsilon_s \text{ and } \epsilon_m$ are the Stommel number and Munk
scale, respectively. The Stommel number and Munk scale are given by (equation
(10) in \cite{Myers})
\begin{equation*}
\epsilon_m = \frac{A}{\beta L^3},
\end{equation*}
where $A$ is the eddy viscosity. The model is supplemented with appropriate
boundary conditions, which will be described for each of the subsequent
numerical tests.
We note that the linear Stommel-Munk model \eqref{eqn:Stommel-Munk} is similar
in form to the SQGE \eqref{eqn:SQGE_Psi}. Indeed, both models contain the
biharmonic operator $\Delta^2 \psi$, the rotation term $\dfrac{\partial
\psi}{\partial x}$, and the forcing terms $f$ and $F$, respectively. The two
main differences between the two models are the following: First, the SQGE are
nonlinear, since they contain the Jacobian term $J(\psi,q)$; the Stommel-Munk
model is linear, since it doesn't contain the Jacobian term. The second
difference is that the linear Stommel-Munk model contains a Laplacian term
$\Delta \psi$, whereas the SQGE does not.
We also note that the two models use different parameters: the Reynolds number
$Re$ and the Rossby number $Ro$ in \eqref{eqn:SQGE_Psi} and the Stommel number
$\epsilon_s$ and the Munk scale $\epsilon_m$ in the linear Stommel-Munk model.
These two sets of parameters, however, are related by the following relations:
\begin{align}
\epsilon_m &= Ro\, Re^{-1}, \label{eqn:Munkscale}\\
\epsilon_s &= Ro \frac{\gamma\, L}{U}. \label{eqn:Stommelnumber}
\end{align}
%\textcolor{red}{If $\gamma \sim T$ then $\epsilon = Ro$. However, I am unsure this is a valid
%assumption and so I am having trouble relating the Stommel number to Reynolds and Rosby.}
The rest of the section is organized as follows: In \autoref{sse:LSM} we
present results for the linear Stommel model \eqref{eqn:Stommel}. In
\autoref{sse:SMM} we present results for the linear Stommel-Munk model
\eqref{eqn:Stommel-Munk}. Finally, in \autoref{sse:SQGE} we present results for
the nonlinear SQGE \eqref{eqn:SQGE_Psi}.
\subsection{Linear Stommel Model} \label{sse:LSM} This subsection presents the
results for the FE discretization of the linear Stommel model
\eqref{eqn:Stommel} by using the Argyris element. The computational domain is
$\Omega = [0,1]\times [0,1]$. For completeness, we present results for two
numerical tests. The first test, denoted by Test 1, corresponds to the exact
solution used by Vallis (equation (14.38) in \cite{Vallis06}), while the second
test, denoted by Test 2, corresponds to the exact solution used by Myers
\emph{et al.} (Equations 15 and 16 in \cite{Myers}).
\tbf{Test 1a:} In this test, we choose the same setting as that used by Vallis
(equation (14.38) in \cite{Vallis06}). In particular, the forcing term and the
non-homogeneous Dirichlet boundary conditions are chosen to match those given by
the exact solution
\begin{equation}
\psi(x,y) = (1-x-e^{-\nicefrac{x}{\epsilon_s}})\, \pi\, \sin \pi y.
\label{eqn:MyerExact}
\end{equation}
We choose the same Stommel number as that used by Vallis, i.e.
$\epsilon_s=0.04$. The exact solution \eqref{eqn:MyerExact} considered by Vallis
satisfies $\psi \to 0$ as $x \to 0$, but does not vanish at $x=1$. In our
numerical tests, we used a standard lifting procedure to treat these
non-homogeneous boundary conditions, i.e. for a problem of the form
\begin{align*}
L \psi&=f \text{ on } \Omega\\
\psi &=g \text{ on } \partial \Omega,
\end{align*}
we reformulate it to be
\begin{align*}
L\tilde{\psi} &= \tilde{f} \text{ on } \Omega \\
\tilde{\psi} &= 0 \text{ on } \partial \Omega
\end{align*}
where $L\tilde{\psi} = Lu - LS = f - g = \tilde{f}$. The solution $\psi$ from
the original problem can then be found by $\psi(x,y) =\tilde{\psi}+S$. The
function $S(x,y)$ is assumed to have the form
\begin{equation*}
S(x,y) = A(y) (1-x) + B(y) x
\end{equation*}
and satisfies the boundary conditions given by \eqref{eqn:MyersExact}. After
some simple algebra we see that
\begin{equation*}
S(x,y) = -x e^{-x/\epsilon_s}\pi \sin(\pi y).
\end{equation*}
The function $\tilde{f}$ can be determined by applying the operator $L$
corresponding to the \emph{linear Stommel} problem to $u - S$.
Applying the finite element method to the \emph{linear Stommel} problem with the
new modified $\tilde{f}$, corresponding to the exact solution given by Vallis,
and homogeneous boundary conditions using Argyris Finite Elements we get a
solution that matches the solution presented by Vallis, as can be seen in
\autoref{fig:StommelVallis}. Additionally, the table of errors
\autoref{tab:StommelErrorsVallis} shows the order of convergence appears to be
approaching the expected rates for $L^2, H^1, \text{ and } H^2$ norms.
\autoref{fig:StommelVallis} presents the streamlines of the approximate solution
obtained by using the Argyris Finite Element on a mesh with $h=\frac{1}{32}$ and
$9670$ DoFs. Comparing \autoref{fig:StommelVallis} with Figure $14.5$ in
\cite{Vallis06}, we notice that our approximation is close to his. Since the
exact solution is available, we can compute the errors in various norms.
\autoref{tab:StommelErrorsVallis} presents the errors $e_0,\, e_1, \text{ and }
e_2$ (i.e., the $L^2,\, H^1, \text{ and } H^2$ errors, respectively) for various
values of the mesh sizes, $h$ (the DoFs are also included).
\begin{table}%[H]
\begin{center}
%{\footnotesize
\begin{tabular}{|c|c|c|c|c|c|c|c|}%c|c|}
\hline
$h$ & $DoFs$ & $e_0$ & $L_2$ order & $e_1$ & $H^1$ order & $e_2$ & $H^2$ order \\[0.2em] % & $e_{\infty}$ & $L_{\infty}$ order \\
\hline
$\nicefrac{1}{2}$ & $70$ & $1.15\times 10^{-1}$ & $-$ & $1.81\times 10^0$ & $-$ & $8.37\times 10^1$ & $-$ \\[0.2em] % & $6.81\times 10^{-3}$ & $-$ \\
$\nicefrac{1}{4}$ & $206$ & $1.02\times 10^{-2}$ & $3.50$ & $3.12\times 10^{-1}$ & $2.54$ & $2.55\times 10^1$ & $1.72$ \\[0.2em] % & $7.801\times 10^{-3}$ & $-1.961\times 10^{-1}$ \\
$\nicefrac{1}{8}$ & $694$ & $4.46\times 10^{-4}$ & $4.51$ & $2.59\times 10^{-2}$ & $3.59$ & $3.90\times 10^0$ & $2.71$ \\[0.2em] % & $0.4.681\times 10^{-2}$ & $4.059$ \\
$\nicefrac{1}{16}$ & $2534$ & $1.09\times 10^{-5}$ & $5.36$ & $1.22\times 10^{-3}$ & $4.41$ & $3.49\times 10^{-1}$ & $3.48$ \\[0.2em] % & $1.732\times 19^{-5}$ & $4.756$ \\
$\nicefrac{1}{32}$ & $9670$ & $1.97\times 10^{-7}$ & $5.79$ & $4.35\times 10^{-5}$ & $4.80$ & $2.34\times 10^{-2}$ & $3.90$ \\[0.2em] % & $5.396\times 19^{-7}$ & $5.005$ \\
\hline
\end{tabular}
%}
\end{center}
\caption{Errors and rate of convergence for the linear Stommel Model
\eqref{eqn:Stommel}, Test 1 \cite{Vallis06}.}
\label{tab:StommelErrorsVallis}
\end{table}
\begin{figure}
\begin{center}
\includegraphics[scale=0.5]{figures/StommelAConvergence.pdf}
\caption{Linear Stommel model \eqref{eqn:Stommel}, Test 1a \cite{Vallis06}:
Observed errors versus expected rate of convergence.}
\label{fig:StommelErrorsVallis}
\end{center}
\end{figure}
\begin{figure}%[H]
\begin{center}
\includegraphics[scale=0.5]{Figures/linearStommelVallis.pdf}
\caption{Linear Stommel model \eqref{eqn:Stommel}, Test 1a \cite{Vallis06}: Streamlines of the approximation,
$\psi^h$, $h=\frac{1}{32}$, and $9670$ DoFs.}
\label{fig:StommelVallis}
\end{center}
\end{figure}
We note that the errors in \autoref{tab:StommelErrorsVallis} follow the
theoretical rates of convergence predicted by the estimates \eqref{eqn:H2Error}
- \eqref{eqn:L2Error} in \autoref{thm:Errors}. The orders of convergence in
\autoref{tab:StommelErrorsVallis} are close to the theoretical ones for the fine
meshes, but not as close for the coarse meshes. We think that the inaccuracies
on the coarse meshes are due to their inability to capture the thin boundary
layer on the left-hand side (i.e., at $x=0$). The finer the mesh gets, the
better this boundary layer is captured and the better the numerical accuracy
becomes.
\tbf{Test 1b:}
In the Second part of Test 1, we verify the hypothesis above, that is, whether
the degrading accuracy of the approximation is indeed due to the thin western
boundary layer. To this end, we change the Stommel number in Test 1a to be
$\epsilon_s=1$, which will result in a much thicker western boundary layer. We
then run the same numerical test as before, but with the new Stommel number. As
can be seen in \autoref{tab:StommelErrorsVallise1}, the rates of convergence are
the expected theoretical orders of convergence. This shows that the reason for
the inaccuracies in \autoref{tab:StommelErrorsVallis} were indeed due to the
thin western boundary layer.
\begin{table}%[H]
\begin{center}
%{\scriptsize
\begin{tabular}{|c|c|c|c|c|c|c|c|}%c|c|}
\hline
$h$ & $DoFs$ & $e_0$ & $L_2$ order & $e_1$ & $H^1$ order & $e_2$ & $H^2$ order \\[0.2em] % & $e_{\infty}$ & $L_{\infty}$ order \\
\hline
$\nicefrac{1}{2}$ & $70$ & $1.69\times 10^{-5}$ & $-$ & $3.43\times 10^{-4}$ & $-$ & $8.72\times 10^{-3}$ & $-$ \\[0.2em] % & $4.306\times 10^{-6}$ & $-$ \\
$\nicefrac{1}{4}$ & $206$ & $3.72\times 10^{-7}$ & $5.50$ & $1.34\times 10^{-5}$ & $4.68$ & $5.62\times 10^{-4}$ & $3.96$ \\[0.2em] % & $5.542\times 10^{-7}$ & $2.958$ \\
$\nicefrac{1}{8}$ & $694$ & $4.89\times 10^{-9}$ & $6.25$ & $3.76\times 10^{-7}$ & $5.16$ & $3.25\times 10^{-5}$ & $4.11$ \\[0.2em] % & $9.043\times 10^{-9}$ & $5.937$ \\
$\nicefrac{1}{16}$ & $2534$ & $7.08\times 10^{-11}$ & $6.11$ & $1.12\times 10^{-8}$ & $5.07$ & $1.96\times 10^{-6}$ & $4.05$ \\[0.2em] % & $1.355\times 10^{-10}$ & $6.06$ \\
$\nicefrac{1}{32}$ & $9670$ & $1.08\times 10^{-12}$ & $6.04$ & $3.44\times 10^{-10}$ & $5.02$ & $1.21\times 10^{-7}$ & $4.02$ \\[0.2em] % & $2.169\times 10^{-12}$ & $5.965$ \\
\hline
\end{tabular}
%}
\end{center}
\caption{Errors and rate of convergence for the linear Stommel Model
\eqref{eqn:Stommel}, Test 1b \cite{Vallis06}.}
\label{tab:StommelErrorsVallise1}
\end{table}
\begin{figure}
\begin{center}
\includegraphics[scale=0.5]{figures/StommelBConvergence.pdf}
\caption{Linear Stommel model \eqref{eqn:Stommel}, Test 1b \cite{Vallis06}:
Observed errors versus expected rate of convergence.}
\label{fig:StommelErrorsVallis1}
\end{center}
\end{figure}
\begin{figure}%[H]
\begin{center}
\includegraphics[scale=0.5]{Figures/StommelVallise1.pdf}
\caption{Linear Stommel model \eqref{eqn:Stommel}, Test 1b \cite{Vallis06}: Streamlines of the approximation,
$\psi^h$, $h=\frac{1}{32}$, and $9670$ DoFs with $\epsilon_s=1$.}
\label{fig:StommelVallise1}
\end{center}
\end{figure}
\tbf{Test 2:}
For our second test we use the exact solution given by Myers (Equations 15 and
16 in \cite{Myers}), i.e.
{\footnotesize
\begin{equation}
\psi(x,y) =\frac{\sin(\pi y)}{\pi(1+4\pi^2\epsilon_s^2)}\left\{2\pi\epsilon_s\sin(\pi x)+cos(\pi x)+\frac{1}{e^{R_1}-e^{R_2}}\left[(1+e^{R_2})e^{R_1x}-(1+e^{R_1})e^{R_2x}\right]\right\},
\label{eqn:MyersExact}
\end{equation}
}
where $R_1\text{ and } R_2$ are the positive and negative roots, respectively,
of
\begin{equation*}
R = \frac{-1\pm\sqrt{1+4\pi^2 \epsilon_s^2}}{2\epsilon_s}.
\end{equation*}
The forcing term and the homogeneous Dirichlet boundary conditions are chosen to
match those given by the exact solution \eqref{eqn:MyersExact}. We choose the
same Stommel number as that used by Myers, i.e. $\epsilon_s=0.05$.
\autoref{fig:StommelMyers} presents the streamlines of the approximate solution
obtained by using the Argyris Finite Element on a mesh with $h=\frac{1}{32}$ and
$9670$ DoFs. Comparing \autoref{fig:StommelMyers} with Figure $2$ in
\cite{Myers}, we notice that our approximation is close to that in \cite{Myers}.
Since the exact solution is available, we can compute the errors in various
norms. \autoref{tab:StommelErrorsMyers} presents the errors $e_0,\, e_1, \text{
and } e_2$ (i.e., the $L^2,\, H^1, \text{ and } H^2$ errors, respectively) for
various values of the mesh sizes, $h$.
\begin{table}%[H]
\begin{center}
%{\footnotesize
\begin{tabular}{|c|c|c|c|c|c|c|c|}%c|c|}
\hline
$h$ & $DoFs$ & $e_0$ & $L_2$ order & $e_1$ & $H^1$ order & $e_2$ & $H^2$ order \\[0.2em] % & $e_{\infty}$ & $L_{\infty}$ order \\
\hline
$\nicefrac{1}{2}$ & $70$ & $5.65\times 10^{-3}$ & $-$ & $1.45\times 10^{-1}$ & $-$ & $6.60\times 10^0$ & $-$ \\[0.2em] % & $8.815\times 10^{-3}$ & $-$ \\
$\nicefrac{1}{4}$ & $206$ & $4.28\times 10^{-4}$ & $3.72$ & $2.08\times 10^{-2}$ & $2.80$ & $1.63\times 10^0$ & $2.02$ \\[0.2em] % & $0.3.139\times 10^{-2}$ & $1.49$ \\
$\nicefrac{1}{8}$ & $694$ & $1.46\times 10^{-5}$ & $4.87$ & $1.41\times 10^{-3}$ & $3.89$ & $2.07\times 10^{-1}$ & $2.98$ \\[0.2em] % & $2.15\times 10^{-5}$ & $3.868$ \\
$\nicefrac{1}{16}$ & $2534$ & $2.95\times 10^{-7}$ & $5.63$ & $5.83\times 10^{-5}$ & $4.59$ & $1.65\times 10^{-2}$ & $3.65$ \\[0.2em] % & $7.097\times 10^{-7}$ & $4.921$ \\
$\nicefrac{1}{32}$ & $9670$ & $4.97\times 10^{-9}$ & $5.89$ & $2.00\times 10^{-6}$ & $4.87$ & $1.07\times 10^{-3}$ & $3.95$ \\[0.2em] % & $2.054\times 10^{-8}$ & $5.111$ \\
\hline
\end{tabular}
%}
\end{center}
\caption{Errors and rate of convergence for the linear Stommel Model
\eqref{eqn:Stommel}, Test 2 \cite{Myers}.}
\label{tab:StommelErrorsMyers}
\end{table}
\begin{figure}
\begin{center}
\includegraphics[scale=0.5]{Figures/StommelMyersConvergence.pdf}
\caption{Linear Stommel model \eqref{eqn:Stommel}, Test 2 \cite{Myers}:
Observed errors versus expected rate of convergence.}
\label{fig:StommelErrorsMeyers}
\end{center}
\end{figure}
\begin{figure}%[H]
\begin{center}
\includegraphics[scale=0.5]{Figures/linearStommelMyers.pdf}
\caption{Linear Stommel model \eqref{eqn:Stommel}, Test 2 \cite{Myers}: Streamlines of the approximation,
$\psi^h$, $h=\frac{1}{32}$, and $9670$ DoFs.}
\label{fig:StommelMyers}
\end{center}
\end{figure}
We note that the errors in \autoref{tab:StommelErrorsMyers} follow the
theoretical rates of convergence predicted by the estimates \eqref{eqn:H2Error}
- \eqref{eqn:L2Error} in \autoref{thm:Errors}. Again, we see that the orders of
convergence in \autoref{tab:StommelErrorsMyers} are close to the theoretical
ones for the fine meshes, but not as close for the coarse meshes. We attribute
this to the inaccuracies at the thin boundary layer on the left-hand side (i.e.,
at $x=0$). The finer the mesh gets, the better this boundary layer is captured
and the better the numerical accuracy becomes.
\subsection{Linear Stommel-Munk Model}\label{sse:SMM}
This subsection presents results for the FE discretization of the linear
Stommel-Munk model \eqref{eqn:Stommel-Munk} by using the Argyris element. Our
computational setting is the same as that used by Cascon \emph{et al.}
\cite{Cascon}: The computational domain is $\Omega = [0,3]\times[0,1]$, the Munk
scale is $\epsilon_m=6\times 10^{-5}$, the Stommel number is $\epsilon_s=0.05$,
and the boundary conditions are
\begin{equation} \label{eqn:SMProb}
\psi = \frac{\partial \psi}{\partial \mathbf{n}}=0 \quad \text{ on } \partial\Omega
\end{equation}
For completeness, we present results for two numerical tests, denoted by Test 3
and Test 4, both corresponding to Test 1 and Test 2 in \cite{Cascon},
respectively.
\tbf{Test 3:}
For our third test we use the exact solution given by Test 1 in \cite{Cascon},
i.e.
\begin{equation}
\psi(x,y) = \sin^2 \frac{\pi x}{3} \sin^2 \pi y.
\label{eqn:CasconExact1}
\end{equation}
The forcing term is chosen to match that given by the exact solution
\eqref{eqn:CasconExact1}.
For this third test we take $F$ corresponding to applying the linear operator
$L$ associated with the \emph{linear Stommel-Munk} model to the exact solution
\eqref{eqn:CasconExact1}.
\autoref{fig:StommelMunkSin} presents the streamlines of the approximate
solution obtained by using the Argyris Finite Element on a mesh with
$h=\frac{1}{32}$ and $28550$ DoFs. Comparing \autoref{fig:StommelMunkSin} with
Figure $7$ in \cite{Cascon}, we notice that our approximation is close to that in
\cite{Myers}. Since the exact solution is available, we can compute the errors
in various norms. \autoref{tab:SMsinErrors} presents the errors $e_0,\, e_1,
\text{ and } e_2$ (i.e., the $L^2,\, H^1, \text{ and } H^2$ errors,
respectively) for various values of the mesh sizes, $h$.
\begin{table}%[H]
\begin{center}
%{\scriptsize
\begin{tabular}{|c|c|c|c|c|c|c|c|}%c|c|}
\hline
$h$ & $DoFs$ & $e_0$ & $L_2$ order & $e_1$ & $H^1$ order & $e_2$ & $H^2$ order \\[0.2em] % & $e_{\infty}$ & $L_{\infty}$ order \\
\hline
$\nicefrac{1}{2}$ & $170$ & $2.99\times 10^{-3}$ & $-$ & $4.08\times 10^{-2}$ & $-$ & $7.62\times 10^{-1}$ & $-$ \\[0.2em] % & $3.423\times 10^{-3}$ & $-$ \\
$\nicefrac{1}{4}$ & $550$ & $3.22\times 10^{-5}$ & $6.54$ & $1.03\times 10^{-3}$ & $5.31$ & $4.08\times 10^{-2}$ & $4.23$ \\[0.2em] % & $1.885\times 10^{-5}$ & $7.505$ \\
$\nicefrac{1}{8}$ & $1958$ & $3.44\times 10^{-7}$ & $6.55$ & $2.49\times 10^{-5}$ & $5.37$ & $2.25\times 10^{-3}$ & $4.18$ \\[0.2em] % & $2.371\times 10^{-7}$ & $6.313$ \\
$\nicefrac{1}{16}$ & $7366$ & $4.57\times 10^{-9}$ & $6.23$ & $7.03\times 10^{-7}$ & $5.15$ & $1.34\times 10^{-4}$ & $4.07$ \\[0.2em] % & $4.296\times 10^{-9}$ & $5.786$ \\
$\nicefrac{1}{32}$ & $28550$ & $6.70\times 10^{-11}$ & $6.09$ & $2.11\times 10^{-8}$ & $5.06$ & $8.26\times 10^{-6}$ & $4.02$ \\[0.2em] % & $6.86\times 10^{-11}$ & $5.969$ \\
\hline
\end{tabular}
%}
\end{center}
\caption{Errors and rate of convergence for the linear Stommel-Munk Model
\eqref{eqn:SMProb}, Test 3 \cite{Cascon}.}
\label{tab:SMsinErrors}
\end{table}
\begin{figure}
\begin{center}
\includegraphics[scale=0.5]{Figures/SMsinConvergence.pdf}
\caption{Linear Stommel-Munk model \eqref{eqn:SMProb}, Test 3 \cite{Cascon}:
Observed errors versus expected rate of convergence.}
\label{fig:SMSinErrors}
\end{center}
\end{figure}
\begin{figure}%[H]
\begin{center}
\includegraphics[scale=0.5]{Figures/StommelMunk1.pdf}
\caption{Linear Stommel-Munk model \eqref{eqn:SMProb}, Test 3 \cite{Cascon}:
Streamlines of the approximation, $\psi^h$, on a mesh size,
$h=\frac{1}{32}$, and $28550$ DoFs.}
\label{fig:StommelMunkSin}
\end{center}
\end{figure}
We note that the errors in \autoref{tab:SMsinErrors} follow the theoretical
rates of convergence predicted by the estimates \eqref{eqn:H2Error} -
\eqref{eqn:L2Error} in \autoref{thm:Errors}. This time, we see that the orders
of convergence in \autoref{tab:SMsinErrors} are close to the theoretical ones
for the fine meshes, but are higher than expected for the for the coarse meshes.
We attribute this to the fact that the exact solution \eqref{eqn:CasconExact1}
does not display any boundary layers that could be challenging to capture by the
Argyris element on a coarse mesh.
\tbf{Test 4:}
For our fourth test we use the exact solution given by Test 2 in \cite{Cascon},
i.e.
{\small
\begin{equation}
\psi(x,y) = \left[\left(1 - \frac{x}{3}\right)\left(1-e^{-20x}\right) \sin \pi y\right]^2.
\label{eqn:CasconExact2}
\end{equation}
}
Again we take the forcing term $F$ corresponding the exact solution
\eqref{eqn:CasconExact2}.
\autoref{fig:SMe} presents the streamlines of the approximate solution obtained
by using the Argyris Finite Element on a mesh with $h=\frac{1}{32}$ and $28550$
DoFs. Comparing \autoref{fig:SMe} with Figure $10$ in \cite{Myers}, we notice
that our approximation is close to \cite{Myers}. Since the exact solution is
available, we can compute the errors in various norms. \autoref{tab:SMeErrors}
presents the errors $e_0,\, e_1, \text{ and } e_2$ (i.e., the $L^2,\, H^1,
\text{ and } H^2$ errors, respectively) for various values of the mesh sizes,
$h$.
\begin{table}%[H]
\begin{center}
%{\small
\begin{tabular}{|c|c|c|c|c|c|c|c|}%c|c|}
\hline
$h$ & $DoFs$ & $e_0$ & $L_2$ order & $e_1$ & $H^1$ order & $e_2$ & $H^2$ order \\[0.2em] % & $e_{\infty}$ & $L_{\infty}$ order \\
\hline
$\nicefrac{1}{2}$ & $170$ & $6.04\times 10^{-2}$ & $-$ & $1.16\times 10^0$ & $-$ & $3.90\times 10^1$ & $-$ \\[0.2em] % & $2.907\times 10^{-2}$ & $-$ \\
$\nicefrac{1}{4}$ & $550$ & $1.13\times 10^{-2}$ & $2.41$ & $4.00\times 10^{-1}$ & $1.54$ & $2.14\times 10^1$ & $0.866$ \\[0.2em] % & $5.678\times 10^{-3}$ & $2.356$ \\
$\nicefrac{1}{8}$ & $1958$ & $8.40\times 10^{-4}$ & $3.75$ & $5.91\times 10^{-2}$ & $2.76$ & $5.66\times 10^0$ & $1.92$ \\[0.2em] % & $0.5.973\times 10^{-2}$ & $3.249$ \\
$\nicefrac{1}{16}$ & $7366$ & $2.82\times 10^{-5}$ & $4.90$ & $4.01\times 10^{-3}$ & $3.88$ & $7.38\times 10^{-1}$ & $2.94$ \\[0.2em] % & $2.979\times 10^{-5}$ & $4.326$ \\
$\nicefrac{1}{32}$ & $28550$ & $5.59\times 10^{-7}$ & $5.66$ & $1.61\times 10^{-4}$ & $4.64$ & $5.97\times 10^{-2}$ & $3.63$ \\[0.2em] % & $8.632\times 10^{-7}$ & $5.109$ \\
\hline
\end{tabular}
%}
\end{center}
\caption{Errors and rate of convergence for the linear Stommel-Munk Model
\eqref{eqn:SMProb}, Test 4 \cite{Cascon}.}
\label{tab:SMeErrors}
\end{table}
\begin{figure}
\begin{center}
\includegraphics[scale=0.5]{Figures/SMeConvergence.pdf}
\caption{Linear Stommel-Munk model \eqref{eqn:SMProb}, Test 4 \cite{Cascon}:
Observed errors versus expected rate of convergence.}
\label{fig:SMeErrors}
\end{center}
\end{figure}
\begin{figure}%[H]
\begin{center}
\includegraphics[scale=0.5]{Figures/StommelMunk2.pdf}
\caption{Linear Stommel-Munk model \eqref{eqn:SMProb}, Test 4 \cite{Cascon}: Streamlines of the approximation,
$\psi^h$, $h=\frac{1}{32}$, and $28550$ DoFs.}
\label{fig:SMe}
\end{center}
\end{figure}
We note that the errors in \autoref{tab:SMeErrors} follow the theoretical rates
of convergence predicted by the estimates \eqref{eqn:H2Error} -
\eqref{eqn:L2Error} in \autoref{thm:Errors}. Again, we see that the orders of
convergence in \autoref{tab:SMeErrors} are close to the theoretical ones for the
fine meshes, but not as close for the coarse meshes. As stated previously, we
attribute this to the inaccuracies at the thin boundary layer on the left-hand
side (i.e., at $x=0$). The finer the mesh gets, the better this boundary layer
is captured and the better the numerical accuracy becomes.
\subsection{SQGE}\label{sse:SQGE}
This subsection presents results for the FE discretization of the streamfunction
formulation of the SQGE \eqref{eqn:SQGE_Psi} by using the Argyris element. Our
computational domain is $\Omega=[0,3]\times[0,1]$, the Reynolds number is
$Re=1.667$, and the Rossby number is $Ro=10^{-4}$. For completeness, we present
results for two numerical tests, denoted by Test 5 and Test 6, both
corresponding to the exact solutions given in Test 1 and Test 2 of
\cite{Cascon}, respectively.
\tbf{Test 5:}
In this test, we take the same exact solution presented in \emph{Test 3}, i.e.
\begin{equation}
\psi(x,y) = \sin^2 \frac{\pi x}{3} \, \sin^2 \pi y.
\label{eqn:StreamfunctionExact1}
\end{equation}
Again, the forcing term $F$ and homogeneous boundary conditions, $\psi = \frac{\partial
\psi}{\partial \mathbf{n}} = 0$, correspond to the exact solution \eqref{eqn:StreamfunctionExact1}.
\autoref{fig:SQGEsin} presents the streamlines of the approximate solution
obtained by using the Argyris Finite Element on a mesh with $h=\frac{1}{32}$ and
$36150$ DoFs. We note that the streamlines look as we expect and are similar to
those given by Figure $7$ in \cite{Myers}, which uses the same exact solution.
Since the exact solution is available, we can compute the errors in various
norms. \autoref{tab:SQGEsinErrors} presents the errors $e_0,\, e_1, \text{ and }
e_2$ (i.e., the $L^2,\, H^1, \text{ and } H^2$ errors, respectively) for various
values of the mesh sizes, $h$.
\begin{table}%[H]
\begin{center}
%{\scriptsize
\begin{tabular}{|c|c|c|c|c|c|c|c|}%c|c|}
\hline
$h$ & $DoFs$ & $e_0$ & $L_2$ order & $e_1$ & $H^1$ order & $e_2$ & $H^2$ order \\[0.2em] % & $e_{\infty}$ & $L_{\infty}$ order \\
\hline
$\nicefrac{1}{2}$ & $138$ & $4.10\times 10^{-3}$ & $-$ & $4.91\times 10^{-2}$ & $-$ & $8.61\times 10^{-1}$ & $-$\\
$\nicefrac{1}{4}$ & $536$ & $2.23\times 10^{-5}$ & $7.52$ & $7.66\times 10^{-4}$ & $6.00$ & $2.96\times 10^{-2}$ & $4.86$\\
$\nicefrac{1}{8}$ & $2349$ & $2.35\times 10^{-7}$ & $6.57$ & $1.60\times 10^{-5}$ & $5.58$ & $1.31\times 10^{-3}$ & $4.50$\\
$\nicefrac{1}{16}$ & $9152$ & $3.03\times 10^{-9}$ & $6.28$ & $4.33\times 10^{-7}$ & $5.21$ & $7.53\times 10^{-5}$ & $4.12$\\
$\nicefrac{1}{32}$ & $36150$ & $4.64\times 10^{-11}$ & $6.03$ & $1.37\times 10^{-8}$ & $4.99$ & $4.84\times 10^{-6}$ & $3.96$\\
$\nicefrac{1}{64}$ & $146090$ & $9.86\times 10^{-13}$ & $5.56$ & $4.10\times
10^{-10}$ & $5.06$ & $2.92\times 10^{-7}$ & $4.05$ \\[0.2em]
\hline
\end{tabular}
%}
\end{center}
\caption{Errors and rate of convergence for the SQGE \eqref{eqn:SQGE_Psi}, Test 5.}
\label{tab:SQGEsinErrors}
\end{table}
\begin{figure}
\begin{center}
\includegraphics[scale=0.5]{Figures/SQGEsin2Convergence.pdf}
\caption{SQGE \eqref{eqn:SQGE_Psi}, Test 5: Observed errors versus expected
rate of convergence.}
\label{fig:SQGEsinErrors}
\end{center}
\end{figure}
\begin{figure}%[H]
\begin{center}
\includegraphics[scale=0.5]{Figures/SQGEsin.pdf}
\caption{SQGE \eqref{eqn:SQGE_Psi}, Test 5: Streamlines of the approximation,
$\psi^h$, $h=\frac{1}{32}$, and $36150$ DoFs.}
\label{fig:SQGEsin}
\end{center}
\end{figure}
We note that the errors in \autoref{tab:SQGEsinErrors} follow the theoretical
rates of convergence predicted by the estimates \eqref{eqn:H2Error} -
\eqref{eqn:L2Error} in \autoref{thm:Errors}. Again, since the exact solution
\eqref{eqn:StreamfunctionExact1} does not display any boundary layers, we see
that the orders of convergence in \autoref{tab:SQGEsinErrors} are close to the
theoretical ones for the fine meshes, but are higher than expected for the for
the coarse meshes. We also note the drop in the rate of convergence in the
$L^2$-norm for $\nicefrac{1}{64}$ is due to reaching the machine epsilon and
thus any rate of convergence for mesh size smaller than $\nicefrac{1}{64}$ is
meaningless.
\tbf{Test 6:}
In this test, we take the same exact solution as used in \emph{Test 4}, i.e.
\begin{equation}
\psi(x,y) = \left[\left(1 - \frac{x}{3}\right)\left(1-e^{-20x}\right) \sin \pi y\right]^2.
\label{eqn:StreamfunctionExact2}
\end{equation}
The forcing term $F$ and the homogeneous boundary conditions correspond to the exact solution
\eqref{eqn:StreamfunctionExact2}.
\autoref{fig:SQGEe} presents the streamlines of the approximate solution
obtained by using the Argyris Finite Element on a mesh with $h=\frac{1}{32}$ and
$36150$ DoFs. We note that the streamlines look as we expect and are similar to
those given by Figure $10$ in \cite{Myers}, which uses the same exact solution.
Since the exact solution is available, we can compute the errors in various
norms. \autoref{tab:SQGEeErrors} presents the errors $e_0,\, e_1, \text{ and }
e_2$ (i.e., the $L^2,\, H^1, \text{ and } H^2$ errors, respectively) for various
values of the mesh sizes, $h$.
\begin{table}%[H]
\begin{center}
%{\small
\begin{tabular}{|c|c|c|c|c|c|c|c|}%c|c|}
\hline
$h$ & $DoFs$ & $e_0$ & $L_2$ order & $e_1$ & $H^1$ order & $e_2$ & $H^2$ order \\[0.2em] % & $e_{\infty}$ & $L_{\infty}$ order \\
\hline
$\nicefrac{1}{2}$ & $138$ & $4.36\times 10^{-1}$ & $-$ & $2.22\times 10^0$ & $-$ & $4.58\times 10^1$ & $-$\\
$\nicefrac{1}{4}$ & $536$ & $7.21\times 10^{-3}$ & $5.92$ & $2.86\times 10^{-1}$ & $2.96$ & $1.65\times 10^1$ & $1.48$\\
$\nicefrac{1}{8}$ & $2349$ & $1.92\times 10^{-3}$ & $1.91$ & $1.07\times 10^{-1}$ & $1.42$ & $8.06\times 10^{0}$ & $1.03$\\
$\nicefrac{1}{16}$ & $9152$ & $2.89\times 10^{-5}$ & $6.05$ & $3.94\times 10^{-3}$ & $4.76$ & $6.84\times 10^{-1}$ & $3.56$\\
$\nicefrac{1}{32}$ & $36150$ & $6.63\times 10^{-7}$ & $5.45$ & $1.75\times 10^{-4}$ & $4.50$ & $5.77\times 10^{-2}$ & $3.57$\\
$\nicefrac{1}{64}$ & $146090$ & $1.37\times 10^{-8}$ & $5.60$ & $6.91\times 10^{-6}$ & $4.66$ & $4.36\times 10^{-3}$ & $3.73$\\
$\nicefrac{1}{128}$ & $585048$ & $1.47\times 10^{-10}$ & $6.55$ & $1.63\times 10^{-7}$ & $5.40$ & $2.28\times 10^{-4}$ & $4.26$ \\
\hline
\end{tabular}
%}
\end{center}
\caption{Errors and rate of convergence for the SQGE \eqref{eqn:SQGE_Psi}, Test 6.}
\label{tab:SQGEeErrors}
\end{table}
\begin{figure}
\begin{center}
\includegraphics[scale=0.5]{Figures/SQGEeConvergence.pdf}
\caption{SQGE \eqref{eqn:SQGE_Psi}, Test 6: Observed errors versus expected
rate of convergence.}
\label{fig:SQGEeErrors}
\end{center}
\end{figure}
\begin{figure}%[H]
\begin{center}
\includegraphics[scale=0.5]{Figures/SQGEe.pdf}
\caption{SQGE \eqref{eqn:SQGE_Psi}, Test 6: Streamlines of the approximation,
$\psi^h$, $h=\frac{1}{32}$, and $36150$ DoFs.}
\label{fig:SQGEe}
\end{center}
\end{figure}
We note that the errors in \autoref{tab:SQGEeErrors} follow the theoretical
rates of convergence predicted by the estimates \eqref{eqn:H2Error} -
\eqref{eqn:L2Error} in \autoref{thm:Errors}. Again for an exact solution which
has a western boundary layer, we see that the orders of convergence in
\autoref{tab:SQGEeErrors} are close to the theoretical ones for the fine meshes,
but not as close for the coarse meshes. We attribute this to the inaccuracies at
the thin boundary layer on the left-hand side (i.e., at $x=0$). The finer the
mesh gets, the better this boundary layer is captured and the better the
numerical accuracy becomes.
\subsection*{Mediterranean Sea}
We have created a FE mesh of the Mediterranean Sea using GMSH \cite{GMSH}. The
coastline data was obtained from GSHHS \cite{GSHHS}. Major islands such as
Corsica, Sardinia, and Sicily were connected to the nearest land mass in order
to ensure a unique streamfunction (see the discussions in
\cite{Gunzburger89,van-Gijzen1998}). Additionally, the Atlantic Ocean was
closed off from the Mediterranean Sea at the Straits of Gibraltar, from the Red
Sea at the Suez Canal, and the Sea of Marmara at the Dardanelles Strait, while
the Gulf of Corinth was treated as land. The resultant FE mesh is displayed in
\autoref{fig:MedMesh}.
\begin{figure}
\begin{center}
\includegraphics[scale=0.4]{Figures/MediterraneanMesh.png}
\caption{Mesh of the Mediterranean Sea created using GMSH \cite{GMSH}. The
mesh size corresponds to $h=\frac{1}{320}$ with $DoF=240,342$.}
\label{fig:MedMesh}
\end{center}
\end{figure}
The following experiment used a forcing function given by $F =
\sin\left(\frac{\pi}{4} y\right)$ which is the same forcing function given by Bryan
\cite{Bryan1963}. Note: Bryan, in fact, used $\sin(\frac{\pi}{2} y)$, but on a
domain with a vertical extent which was of length two. In our case the vertical
extent of the Northern Hemisphere is of length one. Thus, the appropriate forcing
function becomes $\sin(\frac{\pi}{4} y)$. We choose similar parameters as those
used in \cite{delSastre04} and are summarized in \autoref{tab:OceanParameters}.
\begin{table}
\begin{center}
\begin{tabular}{|l|l|}
\hline
$A$ & $2000\,m^2s^{-1}$\\
\hline
$\theta_0$ & $40^\circ$ \\
\hline
$\omega$ & $7.2526\times 10^{-5}\,s^{-1}$ \\
\hline
$H$ & $1000\,m$ \\
\hline
$L$ & $1000\,km$ \\
\hline
$r_e$ & $6378.1\,km$ \\
\hline
$\rho$ & $1024\, \nicefrac{kg}{m^3}$ \\
\hline
\end{tabular}
\end{center}
\caption{Table of parameter values used for the simulations of the
Mediterranean Sea \cite{delSastre04}, where $A,\theta_0,\omega,H,L,r_e,\rho$ are the
eddy viscosity, reference angle for the $\beta$-plane approximation, angular
velocity of the Earth, domain height, domain length, radius of the Earth, and
density of seawater, respectively.}
\label{tab:OceanParameters}
\end{table}
Taking the derivative of relation \eqref{eqn:beta_plane} with respect to $y$
gives the following relation for $\beta$ to $f$ and using the equation (2.80)
from \cite{Vallis06} gives
\begin{equation}
\beta = \frac{2\omega}{r_e}\cos \theta_0.
\label{eqn:Beta}
\end{equation}
From the parameters given in \autoref{tab:OceanParameters} we see that
\begin{equation*}
\beta \approx 1.742\times 10^{-11}\, m^{-1}\,s^{-1}.
\end{equation*}
Using this approximation for $\beta$ and \eqref{eqn:velocity_scale} with
$\tau_0 = 0.6\, dyne\, cm^{-2}$ \cite{Hellerman} gives the following
approximation for the characteristic velocity:
\begin{equation*}
% \begin{split}
% U &= \frac{3.1415 \, 0.6 dyne\, cm^{-2}}{1027 \nicefrac{kg}{m^3} \,
% 1000\, m \, 1.742 \times 10^{-11}\,m^{-1}\,s^{-1} \, 1000\, km} \\
U \approx 1.054\times 10^{-2} \nicefrac{m}{s}.
% \end{split}
\end{equation*}
Therefore by \eqref{eqn:rossby_number}, the Rossby number is
\begin{equation*}
% Ro = \frac{1.054\times 10^{-2} \nicefrac{m}{s}}{1.742\times 10^{-11} m^{-1}
% s^{-1} (1000 km)^2}
Ro = 6.051\times 10^{-4}
\end{equation*}
and by \eqref{eqn:reynolds_number} the Reynolds number is
\begin{equation*}
Re = 5.27.
\end{equation*}
The solution obtained by applying the SQGE to the Mediterranean with parameters
given in \autoref{tab:OceanParameters} and forcing function $F =
\sin(\frac{\pi}{2} y)$ can be seen in \autoref{fig:SQGEMed}. The shape of the
streamfunction is very much in agreement with the solution obtained by Galan del
Sastre (Figure 2.22 in \cite{Galan-del-Sastre2004}). Additionally, we note that
the resultant large scale structures in \autoref{fig:SQGEMed} are very similar
to those in \autoref{fig:GyresMed}. It should be noted that some of the
differences between our solution and the solution obtained by Galan del Sastre's
and our solution and the observed large scale structures seen in
\autoref{fig:GyresMed} are likely a result of his inclusion of more islands and
not connected those islands to Europe or Africa. The rates of convergence are
much lower than expected, however this likely is due to the fact that the domain
is non-smooth which is likely causing the solution to be in a space which is
less regular than the $H^6$ regularity \cite{Soane2010} assumed by the error
analysis in \autoref{sec:SQGEErrors}.
\begin{table}
\begin{tabular}{|c|c|c|c|c|c|c|c|}%c|c|}
\hline
$h$ & $DoFs$ & $e_0$ & $L_2$ order & $e_1$ & $H^1$ order & $e_2$ & $H^2$ order \\[0.2em]
\hline
$\nicefrac{1}{20}$ & $1122$ & $2.08\times 10^{-6}$ & $-$ & $1.95\times 10^{-4}$ & $-$ & $4.50\times 10^{-2}$ & $-$ \\
$\nicefrac{1}{40}$ & $4092$ & $8.00\times 10^{-7}$ & $1.38$ & $6.68\times 10^{-5}$ & $1.54$ & $2.50\times 10^{-2}$ & $0.850$ \\
$\nicefrac{1}{80}$ & $15594$ & $2.91\times 10^{-7}$ & $1.46$ & $2.47\times 10^{-5}$ & $1.43$ & $1.49\times 10^{-2}$ & $0.741$ \\
$\nicefrac{1}{160}$ & $60846$ & $1.04\times 10^{-7}$ & $1.49$ & $9.05\times 10^{-6}$ & $1.45$ & $8.67\times 10^{-3}$ & $0.785$ \\
$\nicefrac{1}{320}$ & $240342$ & $3.10\times 10^{-8}$ & $1.75$ & $2.75\times 10^{-6}$ & $1.72$ & $4.35\times 10^{-3}$ & $0.994$ \\
\hline
\end{tabular}
\caption{Observed rates of convergence for SQGE applied to the Mediterranean Sea
with forcing function $F = \sin(\frac{\pi}{2} y)$ and true solution obtained
from fine mesh with $h=\dfrac{1}{640}$.}
\label{tab:SQGEMedConvergence}
\end{table}
\begin{figure}
\begin{subfigure}[b]{\textwidth}
\begin{center}
\includegraphics{Figures/SQGEMedRe5_27Ro6_051E-4h320sin0_25piy.pdf}
\caption{``True'' solution of SQGE on the Mediterranean Sea with $Re = 5.27, Ro
= 6.051\times 10^{-4}, F = \sin( \frac{\pi}{2} y)$, and $955,302$ DoFs
corresponding to $h = \frac{1}{640}$.}
\label{fig:SQGEMed}
\end{center}
\end{subfigure}
\begin{subfigure}[b]{\textwidth}
\begin{center}
\includegraphics[scale=0.5]{Figures/GyresOfMediterranean.pdf}
\caption{Large scale currents of the Mediterranean Sea \cite{Ayoub1998}.}
\label{fig:GyresMed}
\end{center}
\end{subfigure}
\caption{Comparison of large scale structures visible in the numerical results
of the SQGE applied to the Mediterranean Sea and the observed large scale
structures of the Mediterranean Sea.}
\label{fig:MedCompare}
\end{figure}