-
Notifications
You must be signed in to change notification settings - Fork 4
/
cvmix_kpp.tex
executable file
·4997 lines (4399 loc) · 225 KB
/
cvmix_kpp.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
766
767
768
769
770
771
772
773
774
775
776
777
778
779
780
781
782
783
784
785
786
787
788
789
790
791
792
793
794
795
796
797
798
799
800
801
802
803
804
805
806
807
808
809
810
811
812
813
814
815
816
817
818
819
820
821
822
823
824
825
826
827
828
829
830
831
832
833
834
835
836
837
838
839
840
841
842
843
844
845
846
847
848
849
850
851
852
853
854
855
856
857
858
859
860
861
862
863
864
865
866
867
868
869
870
871
872
873
874
875
876
877
878
879
880
881
882
883
884
885
886
887
888
889
890
891
892
893
894
895
896
897
898
899
900
901
902
903
904
905
906
907
908
909
910
911
912
913
914
915
916
917
918
919
920
921
922
923
924
925
926
927
928
929
930
931
932
933
934
935
936
937
938
939
940
941
942
943
944
945
946
947
948
949
950
951
952
953
954
955
956
957
958
959
960
961
962
963
964
965
966
967
968
969
970
971
972
973
974
975
976
977
978
979
980
981
982
983
984
985
986
987
988
989
990
991
992
993
994
995
996
997
998
999
1000
\chapter{\scshape KPP surface ocean boundary layer}
\label{chapter:cvmix_kpp}
\minitoc
\vspace{.5cm}
\begin{mdframed}[backgroundcolor=lightgray!50]
We summarize the KPP surface boundary layer scheme \citep{LargeKPP}
as implemented in CVMix. CVMix provides the same features as
\cite{LargeKPP} and Appendix A of \cite{Dana_etal2006}. For those
not intent on recovering older results, we make available a
simplified treatment of the non-dimensional shape or structure
function $G(\sigma)$. The following CVMix Fortran module is
directly connected to the material in this chapter:
\begin{align*}
& {\tt cvmix\_kpp.F90}
\end{align*}
\end{mdframed}
\section{Elements of the K-profile parameterization (KPP)}
\label{sec:implementation}
The ocean surface boundary layer (OBL) mediates the exchange of
properties between the ocean and other components of the climate
system. Hence, parameterization of processes active in the OBL are
fundamental to the integrity of a climate simulation. The K-profile
parameterization (KPP) is a widely used method for parameterizing
boundary layer processes in both the atmosphere and ocean.\footnote{We
consider here the implementation of KPP for the surface ocean
boundary layer, as implementations for the bottom do not exist in
MOM or POP. \cite{Durski_etal2004} consider KPP for the ocean
bottom in ROMS.} The paper by \cite{LargeKPP} introduced this scheme
to the ocean community for use in parameterizing processes in the
surface ocean boundary layer . The pedagogical lectures by
\cite{LargeKPP_lectures} and \cite{Large2012} provide added insight
into the scheme that complements some of the material in
\cite{LargeKPP}.
The KPP scheme has been used by many ocean climate studies for
parameterizing mixing in the OBL, with examples discussed in
\cite{Largeforcing}, \cite{HollandChowBryan1998},
\cite{Gent_etal_1998}, \cite{GOTM}, \cite{LiChowMcWilliamsFu2001},
\cite{Smyth_etal2002}, \cite{Durski_etal2004}, and
\cite{Chang_etal2005}. It was also used in various climate and earth
system models developed at NCAR and GFDL, and elsewhere.
We aim to thoroughly expose physical aspects of the KPP scheme,
building on discussions in \cite{LargeKPP} and \cite{Large2012}. We
also discuss issues that arose when testing the scheme for CVMix.
These tests motivate suggestions for how to make use of the CVMix
version of KPP. Correspondingly, we propose simplifications available
in CVMix that aim to resolve problems with the \cite{LargeKPP}
implementation.
\subsection{Conventions}
\label{subsection:conventions}
We use the following notational and sign conventions in this chapter.
\begin{itemize}
\item The fluid is assumed to be volume conserving Boussinesq.
Extensions to a mass conserving non-Boussinesq fluid are trivial.
\item \label{geopotential_defined} The geopotential coordinate, $z$,
increases up, with $z=0$ defining the resting ocean surface. The
ocean free surface is at $z=\eta(x,y,t)$ and the static ocean bottom
is at $z=-H(x,y)$. Neither $\eta$ nor $H$ contain overturns.
\item \label{height_defined} The height coordinate, ${\cal H}$, is
defined as minus the depth, $d$, according to the discussion in
Section \ref{section:vertical-grid-numerics}, in which
\begin{equation}
{\cal H} = -d = z - \eta,
\label{eq:height-defined-kpp}
\end{equation}
which lives within the range
\begin{equation}
-(H+\eta) \le {\cal H} \le 0.
\label{eq:height-range-defined-kpp}
\end{equation}
The CVMix version of the KPP code assumes the calling models (e.g.,
MOM, MPAS, POP) provide the height as defined here, so that the CVMix
code does not need to be given the sea level field.
\item \label{lambda_defined} A lowercase $\lambda$ is used to denote a
turbulent fluctuation of an arbitrary field within the surface ocean
boundary layer; e.g., a tracer such as potential or conservative
temperature $\theta$ and salinity $s$, or a velocity component
($u,v,w$). Note that $x$ is the notation used in \cite{LargeKPP} and
\cite{LargeKPP_lectures}, but we prefer the Greek letter $\lambda$
to avoid confusion with the horizontal spatial coordinate. Also,
the symbol $s$ is somtimes used for scalar fields such as salinity
and temperature, whereas $m$ is sometimes used for components of a
vector field.
\item There is no distinction in the treatment of scalar fields within
the KPP boundary layer. It is only beneath the boundary layer,
where double diffusive processes are relevant, that we distinguish
the mixing between scalar fields such as temperature and salinity.
Also, surface forcing for salt/scalars and heat distinguish the
tracers.
\item \label{Lambda_defined} An uppercase $\Lambda$ is used to denote
the Eulerian mean of a tracer or velocity component within the
surface ocean boundary layer; e.g., potential or conservative
temperature $\Theta$, salinity $S$, or velocity component ($U,V,W$).
The Eulerian mean fields are time stepped by an ocean climate model
within the boundary layer, and correlations of turbulent variables
must be parameterized to close the mean field equations.
\item \label{correlation_defined} The expression $\overline{w \,
\lambda}$ is used to symbolize the Eulerian correlation of the
fluctuating turbulent vertical velocity and a fluctuating scalar or
vector field. This correlation appears in the mean field time
tendency equation for $\Lambda$ in the Boussinesq primitive ocean
equations (see equation (\ref{eq:mean-field-equation-kpp})). KPP
provides a parameterization of this vertical turbulent flux within
the surface ocean boundary layer.
\item \label{w_W_defined} The mean and turbulent vertical velocity
components, $W,w$, are positive for upward motion. This sign
convention implies that
\begin{mdframed}[backgroundcolor=lightgray!50]
\begin{equation}
\overline{w \, \lambda} > 0 \implies \mbox{turbulent flux for $\lambda$ transported vertically upward}.
\label{eq:correlation-convention}
\end{equation}
\end{mdframed}
If $\lambda$ is the temperature, then a positive correlation at the
ocean surface,
\begin{equation}
\overline{w \, \theta}^{d=0} > 0,
\end{equation}
corresponds to surface cooling. To reduce notation clutter,
correlations evaluated at the ocean surface will be written
\begin{equation}
\overline{w \, \theta}^{0} = \overline{w \, \theta}^{d=0}.
\end{equation}
\item Boundary fluxes of scalar fields are denoted by a capital $Q$,
along with a subscript or superscript to denote the particular flux.
Such scalar fluxes are assumed to be positive when entering the
ocean and negative when leaving the ocean
\begin{mdframed}[backgroundcolor=lightgray!50]
\begin{equation}
Q > 0 \Rightarrow \mbox{boundary scalar flux enters the ocean.}
\label{eq:boundary-sign-convention}
\end{equation}
\end{mdframed}
\item We make the following observations about the sign conventions
(\ref{eq:correlation-convention}) and
(\ref{eq:boundary-sign-convention}).
\begin{itemize}
\item A positive heat flux, $Q^{\mbox{\footnotesize heat}} > 0$,
either through the ocean surface or ocean bottom, adds heat to the
ocean; likewise for salt and water.
\item The sign convention (\ref{eq:boundary-sign-convention}) is
followed in MOM and POP. However, it is not the convention used in
\cite{LargeKPP}, whose convention was in fact opposite for some
cases except for penetrative radiation.
\item We consider a positive surface buoyancy forcing, $B_{f} > 0$
(units $\mbox{m}^{2}~\mbox{s}^{-3}$), to increase the ocean
buoyancy. Adding heat to the ocean increases its buoyancy in
regions of positive thermal expansion, whereas adding salt decreases
buoyancy in regions of positive haline contraction.
\item The convention (\ref{eq:boundary-sign-convention}) necessitates
a minus sign when equating surface boundary fluxes of scalars to the
correlations $\overline{w \, \lambda}^{0}$ defined by equation
(\ref{eq:correlation-convention}).
\item At the ocean bottom, the convention requires no minus sign,
since $\overline{w \, \lambda}^{d=H+\eta} > 0$ means there is a transfer
of scalar field into the ocean through the ocean bottom, such as
through geothermal heating. Note that we are not concerned with
implementing KPP at the ocean bottom \citep{Durski_etal2004}.
\end{itemize}
\item Momentum imparted to the ocean surface by a boundary stress,
$\bftau$, acts to accelerate the ocean in the respective direction.
In contrast, a positive sign to a component of $\overline{w \, {\bf
u}}^{0}$ removes the associated momentum from the surface
ocean. These sign conventions give rise to the minus sign in the
relation (\ref{eq:wu-kinematic-flux-kpp}) connecting turbulent
kinematic stress to the boundary stress:
\begin{equation}
\overline{w \, {\bf u}}^{0} = -\rho^{-1} \, \bftau.
\end{equation}
\end{itemize}
\subsection{General form of the KPP parameterization}
Ignoring all terms except vertical advective transport in the
prognostic equation for the mean field $\Lambda$, its time tendency
is determined by
\begin{equation}
\frac{\partial \Lambda}{\partial t} = -\left( \frac{\partial \, (W \, \Lambda)}{\partial z} \right)
-\left( \frac{\partial \, (\overline{w \, \lambda}) }{\partial z} \right).
\label{eq:mean-field-equation-kpp}
\end{equation}
The advective flux by the mean vertical velocity, $W \, \Lambda$, is
represented via a numerical advection operator. In contrast, the
turbulent correlation, $\overline{w \, \lambda}$, is a subgrid scale
flux that must be parameterized in order to close the equation for
$\Lambda$. Here, the overbar signifies an Eulerian averaging operator
over unresolved turbulent motions occurring within the OBL.
The KPP scheme provides a first order closure for $\overline{w \,
\lambda}$ within the OBL. It does so by introducing two terms in the
following manner
\begin{equation}
\overline{w \, \lambda} = -K_{\lambda} \left( \frac{\partial \Lambda}{\partial z} \right)
+ K^{\mbox{\tiny non-local}}_{\lambda} \, \gamma_{\lambda}.
\label{eq:kpp-parameterization}
\end{equation}
The KPP prescription (\ref{eq:kpp-parameterization}) thus
parameterizes the vertical turbulent flux according to
\begin{equation}
\overline{w \, \lambda} = \overline{w \, \lambda}^{\mbox{\tiny local}} + \overline{w \, \lambda}^{\mbox{\tiny non-local}}.
\label{eq:vertical-flux-decomposed}
\end{equation}
The first term provides for the familiar downgradient vertical
diffusion determined by a vertical diffusivity and the local vertical
derivative of the mean field. This term is referred to as the local
portion of the parameterization
\begin{equation}
\overline{w \, \lambda}^{\mbox{\tiny local}} = -K_{\lambda} \left( \frac{\partial \Lambda}{\partial z} \right).
\label{eq:vertical-flux-local}
\end{equation}
Note that the diffusivity $K_{\lambda} $ computed from KPP is a
non-local function of boundary layer properties, so the name ``local''
is not directed at the diffusivity, but instead at the vertical
derivative. The second term, $\gamma_{\lambda}$, accounts for
non-local transport that is not directly associated with local
vertical gradients of $\Lambda$, in which
\begin{equation}
\overline{w \, \lambda}^{\mbox{\tiny non-local}} = K^{\mbox{\tiny non-local}}_{\lambda} \; \gamma_{\lambda}.
\label{eq:vertical-flux-nonlocal}
\end{equation}
\subsection{The vertical diffusivity}
\label{subsection:kpp-vertical-diffusivity}
The KPP vertical diffusivity used to parameterize the local flux
(\ref{eq:vertical-flux-local}) in the OBL is determined as a non-local
function of boundary layer properties. It is written in the following
form
\begin{equation}
K_{\lambda}(\sigma) = h \, w_{\lambda}(\sigma) \, G_{\lambda}(\sigma).
\label{eq:kpp-diffusivity}
\end{equation}
The diffusivity is constructed as the product of three terms:
\begin{itemize}
\item boundary layer thickness $h$,
\item vertical turbulent velocity scale $w_{\lambda}(\sigma)$,
\item dimensionless vertical shape or structure function
$G_{\lambda}(\sigma)$.
\end{itemize}
Note that we introduce a dependence of the shape function on the field
diffused. Such dependence can arise if taking the approach of
\cite{LargeKPP} whereby the boundary layer diffusivity is matched at
the base of the boundary layer to the interior diffusivities, which
can generally be a function of the tracer, $\lambda$. However, as
discussed in Section \ref{subsection:kpp-shape-function-cvmix}, the
recommended approach for CVMix is to use the following universal shape
function for all tracers (equation
(\ref{eq:universal-non-local-structure}))
\begin{equation}
G(\sigma)_{\mbox{\tiny universal}} = \sigma \, \ (1-\sigma)^{2},
\label{eq:simpler-shape-function-first}
\end{equation}
thus greatly simplifying the KPP scheme.
\cite{LargeKPP} proposed to set the diffusivities equal
\begin{equation}
K^{\mbox{\tiny non-local}}_{\lambda} = K_{\lambda},
\end{equation}
and we support that recommendation, along with setting the shape
function to the universal form
(\ref{eq:simpler-shape-function-first}). However, in moving from the
\cite{LargeKPP} version of KPP to this simplified version, one may
choose to test an intermediate version in which the non-dimensional
shape function is distinct for $K^{\mbox{\tiny non-local}}_{\lambda}$
and $K_{\lambda}$. It is for this reason that we maintain the
distinct symbols for the diffusivities, even though we recommend users
choose the CVMix implementation of KPP that sets them equal. We have
more to say on this topic in Section
\ref{subsection:kpp-shape-function}.
\subsubsection{Boundary layer thickness}
The boundary layer thickness is denoted by
\begin{equation}
h \ge 0 \; \; \mbox{is the boundary layer thickness}.
\label{eq:boundary-layer-thickness}
\end{equation}
This is the thickness of the OBL prescribed by the KPP scheme, with
details given in Section \ref{subsection:kpp-obl-thickness}. The
surface boundary layer generally thickens when mechanical forcing
mixes the water, and negative buoyancy forcing makes the water
gravitationally unstable. Conversely, the bounday layer shoals with
weak winds and/or positive buoyancy forcing. The direct dependence of
the vertical diffusivity in equation (\ref{eq:kpp-diffusivity}) on the
OBL thickness manifests the common property of boundary layers,
whereby thicker layers generally arise from stronger eddy motions and
are thus associated with more rapid mixing of tracer concentration and
momentum.
Figure \ref{fig:boundary-layer-schematic-kpp} provides a schematic of
the KPP boundary layer, the Monin-Obukhov surface layer, and the
associated momentum, mass, and buoyancy fluxes impacting these layers.
Details of this figure will be explored in the following.
\subsubsection{Measuring vertical distances within the OBL}
When measuring distances within the boundary layer, it is the
thickness of the water as measured from the ocean surface that is
important. Free surface undulations can be a nontrivial fraction of
the boundary layer thickness, particularly under conditions of stable
buoyancy forcing. Hence, we make explicit note that the ocean has an
undulating free surface at $z=\eta(x,y,t)$, which contrasts to
\cite{LargeKPP} and \cite{LargeKPP_lectures}, who assumed that $z=0$
sets the upper ocean surface.
Following \cite{LargeKPP}, we introduce the non-dimensional depth,
$\sigma$, given by
\begin{equation}
\sigma = \frac{d}{h}.
\label{eq:sigma-defined}
\end{equation}
In this definition, $d \ge 0$ is the distance from the ocean surface
at $z=\eta$ to a point within the boundary layer
\begin{equation}
d = -z+\eta.
\label{eq:distance-from-surface-defined}
\end{equation}
Likewise, $h \ge 0$ is the distance from the free surface
to the bottom of the boundary layer
\begin{equation}
h = h_{\mbox{\tiny obl}} +\eta,
\label{eq:h-obl-defined}
\end{equation}
where $h_{\mbox{\tiny obl}}$ is the depth of the boundary layer as
measured from $z=0$. That is, $h$ is the thickness of the OBL, and it is this
thickness, not $h_{\mbox{\tiny obl}}$, that is
predicted by KPP (Section \ref{subsection:kpp-obl-thickness}).
Regions within the boundary layer are given by the non-dimensional
depth range
\begin{equation}
0 \le \sigma \le 1 \qquad \mbox{within boundary layer,}
\end{equation}
with $\sigma=0$ the ocean surface and $\sigma = 1$ the bottom of the
boundary layer.
%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%
\begin{figure}[h!t]
\rule{\textwidth}{0.005in}
\begin{center}
\includegraphics[angle=0,width=10cm]{./mfpic_figs/cvmix_kpp_boundary_layer.pdf}
\caption[KPP boundary layer schematic]{\sf Schematic of the upper
ocean boundary layer regions associated with the KPP boundary layer
parameterization. The upper ocean is exposed to non-penetrative
air-sea and ice-sea fluxes of momentum $\bftau$ (Section
\ref{section:boundary-forcing-momentum-kpp}), mass $\Qm$(Section
\ref{section:boundary-forcing-buoyancy-kpp}), and buoyancy $B_{f}$
(Section \ref{section:boundary-forcing-buoyancy-kpp}). In addition,
there is penetrative shortwave radiation, $-\overline{w \,
\theta}_{R}$ (Section
\ref{section:boundary-forcing-buoyancy-kpp}), indicated by the
exponentially decaying vertical sinusoidal. The Monin-Obukhov
surface layer (Section \ref{section:m-o-similarity}) has a thickness
$\epsilon \, h$, with $\epsilon \approx 0.1$. The surface layer is
where turbulence delivers fluxes to the molecular skin layer for
transfer to the atmosphere or ice. The surface layer starts from
just beneath the surface roughness elements at the upper ocean
interface. Since neither these roughness elements, nor the
molecular viscous sublayer, are resolved in ocean models, we assume
in practice that the Monin-Obukhov surface layer extends to the sea
surface at $z=\eta(x,y,t)$. The KPP boundary layer includes the
surface layer, and it has a thickness $h(x,y,t)$ determined by the
KPP parameterization (Section \ref{subsection:kpp-obl-thickness}).
The ocean bottom at $z=-H(x,y)$ is rigid and is exposed to
geothermal heating. Presently, the KPP boundary layer scheme has
not been implemented in MOM or POP to parameterize bottom boundary
layer physics, though nothing fundamental precludes such. In fact,
\cite{Durski_etal2004} provide just such an implementation.}
\label{fig:boundary-layer-schematic-kpp}
\end{center}
\rule{\textwidth}{0.005in}
\end{figure}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\subsubsection{Scale for turbulent vertical velocity fluctuations $w_{\lambda}$}
\label{subsubsection:turbulent-vertical-velocity-scale}
We introduce a scale for turbulent vertical velocity fluctuations,
written as $w_{\lambda}(\sigma)$. This scale is a function of depth
within the boundary layer, and a function of the field to which it
refers. \cite{LargeKPP} recommend using the {\it same} scale $w_{s}$
for all scalar fields (temperature, salinity, and passive tracers)
\begin{equation}
w_{s}= \mbox{same for all scalars}.
\end{equation}
The scale $w_{s}$ also is the same as the turbulent velocity scale for
momentum, $w_{m}$, in cases where the surface buoyancy forcing,
$B_{f}$ is positive. However, $w_{m} < w_{s}$ under unstable surface
buoyancy forcing
\begin{subequations}
\begin{align}
w_{m} &= w_{s} \qquad B_{f} > 0 \\
w_{m} &< w_{s} \qquad B_{f} < 0.
\end{align}
\end{subequations}
That is, gravitational instability is assumed to mix scalars more
efficiently than momentum. We return to the specification of
$w_{\lambda}$ in Section \ref{subsection:vertical-velocity-scale}.
\subsubsection{Non-dimensional vertical shape function $G_{\lambda}(\sigma)$}
Non-dimensional vertical shape function $G_{\lambda}(\sigma)$ is used to
smoothly transition from the ocean surface to the bottom of the
boundary layer. \cite{LargeKPP} chose a cubic polynomial
\begin{equation}
G_{\lambda}(\sigma) = a_{0} + a_{1} \, \sigma + a_{2} \, \sigma^{2} + a_{3} \, \sigma^{3}.
\label{eq:shape-function-gsigma}
\end{equation}
Since turbulent eddies do not cross the ocean surface at $\sigma=0$,
we should correspondingly have a vanishing diffusivity at $\sigma=0$.
This constraint is satisfied by setting
\begin{equation}
a_{0} = 0.
\end{equation}
We detail in Section \ref{subsection:kpp-shape-function} how to
specify the remaining expansion coefficients $a_{1}, a_{2}, a_{3}$
following the approach from \cite{LargeKPP}, in which the polynomial
coefficients $a_{1}, a_{2}, a_{3}$ depend on the tracer fields in so
far as they are specified by matching to the interior diffusivities.
Double diffusive processes lead to distinct diffusivities for
temperature and material tracers such as salt.
However, as discussed in Section \ref{subsection:kpp-shape-function},
the recommended approach for CVMix is to use the following universal
shape function for all tracers (equation
(\ref{eq:universal-non-local-structure}))
\begin{equation}
G(\sigma)_{\mbox{\tiny universal}} = \sigma \, \ (1-\sigma)^{2},
\end{equation}
thus greatly simplifying the KPP scheme. This universal shape
function satisfies
\begin{equation}
G(0)_{\mbox{\tiny universal}}= G(1)_{\mbox{\tiny universal}} = G'(1)_{\mbox{\tiny universal}} = 0,
\end{equation}
which is the form suggested by atmospheric boundary layer
implementations of KPP (see Figure \ref{fig:kpp-figure2-reproduced}).
\subsection{The non-local transport $\gamma_{\lambda}$}
\label{subsection:kpp-nonlocal-transport-outline}
There are many processes in the boundary layer that lead to transport
that is difficult to parameterize as a function of the local vertical
derivative of the mean field (see Section 2 of \cite{LargeKPP}). This
behaviour leads to a diffusivity $K_{\lambda}$ that is a function of
the surface fluxes and boundary layer thickness $h$. Furthermore,
under convective forcing (negative surface buoyancy forcing; $B_{f} <
0$), fluxes can penetrate into the stratified interior. This
characteristic then motivates the introduction of a non-local
transport term $\gamma_{\lambda}$ to the KPP parameterization
(equation (\ref{eq:kpp-parameterization})) in the case of $B_{f} < 0$.
To further identify the need for a non-local transport term
$\gamma_{\lambda}$, we reproduce Figure 1 from \cite{LargeKPP}, here
shown as Figure \ref{fig:kpp-figure1-reproduced}. The caption to
Figure \ref{fig:kpp-figure1-reproduced} explores the many facets of
this figure used to help justify the non-local term in KPP.
%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%
\begin{figure}[h!t]
\rule{\textwidth}{0.005in}
\begin{center}
\includegraphics[angle=0,width=10cm]{./figs/LargeKPP_fig1.pdf}
\caption[Figure 1 from \cite{LargeKPP}]{
\sf
This is a reproduction of
Figure 1 from \cite{LargeKPP}.
The figure is derived from a one-dimensional simulation after 3 days of
convective deepening (zero winds; negative surface buoyancy forcing)
into an initially uniformly stratified water column. The vertical axis
is vertical distance starting from the ocean surface interface at
$z=\eta$ and $d=0$, extending down to $d=h$ ($h=13.6$~m at this point
of the integration), which is the base of
the boundary layer, and finally to $d=1.4\, h$, which is beneath the
boundary layer.
\hspace{0.4cm} The horizontal axis on the bottom is the mean buoyancy, $B$,
relative to that at the surface, $B_{0}$, and the profile is
depicted by the solid line. Positive values of
$B-B_{0}$ indicate that the mean buoyancy at a point is larger than
at the surface, with $B-B_{0} > 0$ expected under
negative buoyancy forcing at the ocean surface.
\hspace{0.4cm} The horizontal axis on
the top is the ratio of the local turbulent buoyancy flux
$\overline{w \, b}$ to the surface turbulent flux $\overline{w \,
b}^{0}$ (denoted $\overline{w \, b}_{0}$ by \cite{LargeKPP}).
The dashed line depicts this ratio. Positive values of $\overline{w
\, b}$ represent upward turbulent buoyancy fluxes; e.g., upward
fluxes of heat (ocean surface cooling) for the case where buoyancy is determined by
temperature, and the thermal expansion coefficient is positive.
\hspace{0.4cm} Positive values for $\overline{w \, b}$ in regions between roughly $0.35 < d
< 0.8$ represent upward turbulent buoyancy fluxes in a region where the mean vertical
gradient of $B$ is nearly zero, thus indicating non-local turbulent transport.
In shallower regions with $d < 0.35$, the mean gradient is negative,
$\partial_{z} B < 0$, and the fluxes are positive, $\overline{w \,
b} > 0$, thus representing downgradient turbulent fluxes.
Likewise, for $d> 0.8$, the turbulent fluxes are downgradient.
\hspace{0.4cm} The mixed layer depth is denoted by $h_{m}$, though
this depth is subject to arbitrary specification of the density
difference. The entrainment depth is $h_{e}$, with this depth taken
where the buoyancy flux reaches a negative extrema. Note that it is
an empirical result that under pure convective forcing ($\bftau =0,
B_{f} < 0$), the turbulent entrainment flux is roughly 20\% of the
surface flux: $\overline{w \, b}^{d=h_{e}} = \beta_{T} \;
\overline{w \, b}^{0}$, where $\beta_{T} = -0.2$. This situation is
depicted in the figure. }
\label{fig:kpp-figure1-reproduced}
\end{center}
\rule{\textwidth}{0.005in}
\end{figure}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
As part of the KPP parameterization, the non-local transport,
$\gamma_{\lambda}$, aims to account for such processes as boundary
layer eddies whose transport may be unrelated to the local vertical
gradient of the mean field, and whose impacts may penetrate within the
stratified ocean interior. In general, \cite{LargeKPP} prescribe the
following characteristics to $\gamma_{\lambda}$.
\begin{itemize}
\item Page 371 of \citep{LargeKPP} notes that there is no theory for
non-local momentum transport, so the non-local transport is assumed
to directly affect only the tracer fields:
\begin{equation}
\gamma_{\lambda} \; \; = \; \;
\left\{
\begin{array}{ll}
0 \; \; &\mbox{if $\lambda = (u,v,w)$ a velocity component}
\\
\ne 0 \; \; &\mbox{nonzero if $\lambda = \theta,s$ or another tracer.}
\end{array}
\right.
\end{equation}
\cite{Smyth_etal2002} consider a non-local term for momentum, thus
motivating further research to see whether it is suitable for climate
modeling. CVMix has not implemented the \cite{Smyth_etal2002} scheme.
However, CVMix has hooks available in the code for a non-local
momentum transport term, thus facilitating further research into this
question.
\item The non-local transport is non-zero only within the OBL:
\begin{equation}
\gamma_{\lambda} \; \; = \; \;
\left\{
\begin{array}{ll}
0 \; \; &\mbox{if $\sigma > 1$}
\\
\ne 0 \; \; &\mbox{if $0 \le \sigma \le 1$.}
\end{array}
\right.
\end{equation}
\item The non-local transport is non-zero only in the presence of
destabilizing negative surface ocean buoyancy flux, whose presence
gives rise to convective mixing:
\begin{equation}
\gamma_{\lambda} \; \; = \; \;
\left\{
\begin{array}{ll}
0 \; \; &\mbox{for positive (stabilizing) surface buoyancy forcing}
\\
\ne 0 \; \; &\mbox{for negative (destabilizing) surface buoyancy forcing.}
\end{array}
\right.
\end{equation}
\item The non-local transport can give rise, under
certain conditions, to either down-gradient or up-gradient
transport of the mean tracer field. Hence, it can either act to
smooth gradients of mean fields (downgradient non-local fluxes) or
enhance gradients (upgradient non-local fluxes).
\end{itemize}
We summarize the KPP parameterization of $\gamma_{\lambda}$ in Section
\ref{subsection:kpp-non-local-transport}. As shown in that section,
the KPP non-local transport acts to redistribute the surface tracer
fluxes throughout the boundary layer.
\subsection{Some caveats regarding KPP}
\label{subsection:caveats-for-kpp}
Subsequent to \cite{LargeKPP}, many ocean modelling groups have made
use of the KPP scheme for ocean climate purposes. One key practical
feature of the scheme is the general ability to use relatively coarse
vertical grid spacing, more coarse than typically suitable for
turbulent kinetic energy methods such as those coded in \cite{GOTM}.
Physically, the non-local transport term available in KPP can capture
features in the boundary layer that are more difficult, if not
impossible, to parameterize using local methods.
Nonetheless, KPP is not a complete scheme nor is it without its
problems. For example, as we note in Section
\ref{section:kernals-of-behaviour}, there are cases where KPP is
clearly performing in an unphysical manner. One key reason the scheme
can be pushed outside of physical relevance concerns the absence of an
energetic basis for KPP. Furthermore, the non-local scheme is not
constrained to satisfy the 2nd Law of Thermodynamics, meaning that it
can produce extrema. We present potential ``fixes'' in Section
\ref{section:kernals-of-behaviour}. However, these fixes are based
largely on {\it ad hoc} considerations. A more fundamental approach
may be to combine the ideas of KPP with those of energetically based
turbulence closure schemes. Such considerations are the topic of
ongoing research.
\section{Surface ocean boundary momentum fluxes}
\label{section:boundary-forcing-momentum-kpp}
In this section and Section
\ref{section:boundary-forcing-buoyancy-kpp}, we present features of
how surface boundary fluxes force the upper ocean, largely following
Appendix A of \cite{LargeKPP}. The aim is to identify how surface
boundary fluxes impact the upper ocean, with this characterization
then used in Section \ref{section:m-o-similarity} to help establish
some basic features of ocean boundary layers. These ideas are then
used in Section \ref{section:specifying-kpp-diffusivity-nonlocal} to
specify the diffusivity and non-local transport from the KPP
parameterization.
Vertical exchange of momentum across the atmosphere-ocean or
sea-ice-ocean boundary occurs largely through turbulent processes.
The resulting horizontal stress vector acting on the ocean, $\bftau$,
is determined through application of a bulk formula (e.g., see
Appendix C of \cite{CORE_NYF} or \cite{LargeYeager2009}). For our
purposes, we assume $\bftau$ is given, thus yielding the ocean
kinematic fluxes associated with the turbulent transport of momentum
across the ocean surface at $d = -z + \eta = 0$
\begin{equation}
-\overline{w \, {\bf u}}^{0}
= \left( \frac{ \bftau }{\rho(\eta)} \right) \approx \left( \frac{ \bftau }{\rho_{o}} \right).
\label{eq:wu-kinematic-flux-kpp}
\end{equation}
In this equation, $\rho(\eta)$ is the surface ocean density, which is
commonly approximated by the constant Boussinesq reference density
$\rho_{o}$. A positive sign on a component of $\bftau$ acts to
accelerate the flow in the respective direction, whereas a positive
sign to a component of $\overline{w \, {\bf u}}^{0}$ removes
momentum from the ocean. These sign conventions give rise to the
minus sign in the relation (\ref{eq:wu-kinematic-flux-kpp}). In
addition to defining the kinematic surface fluxes, knowledge of
$\bftau$ allows us to compute surface boundary layer velocity scales
when working within the Monin-Obukhov similarity theory (Section
\ref{subsection:m-o-similarity-theory}).
In addition to turbulent momentum transfer, $\bftau$ is associated
with momentum transported through mass exchange across the ocean
surface, since water transported across the ocean generally carries a
nonzero momentum. \cite{KanthaClaysonII} (see their page 431) point
out that this effect can be nontrivial, particularly when resolving
strong atmospheric storms. They also make the case for including this
effect in computing the Monin-Obukhov length scale defined by equation
(\ref{eq:m-o-length-scale}) (see their equation (4.3.11)). Notably,
when running a coupled model, the stress from rain is included, since
it is part of the momentum convergence acting at the bottom of the
atmospheric column. Modifying the stress from a prescribed
atmospheric state, such as CORE \citep{LargeYeager2009}, requires
further considerations.
\section{Surface ocean boundary buoyancy fluxes}
\label{section:boundary-forcing-buoyancy-kpp}
Turbulent and advective fluxes of momentum and buoyancy are
transferred across the upper ocean surface boundary, with ocean
processes such as advection and mixing then transporting the boundary
momentum and buoyancy laterally as well as into the ocean interior.
In contrast, penetrative shortwave radiation is absorbed into the
ocean absent ocean transport processes, with such absorption a
function of ocean optical properties. In the unphysical case of
perfectly transparent seawater, shortwave radiation penetrates through
the boundary layer and so has no influence on boundary layer
processes. In realistic cases, much of the shortwave radiation is
absorbed in the boundary layer, with only a fraction leaking through
to the interior. In general, such non-turbulent and non-advective
transport of buoyancy via penetrative radiation represents a
fundamentally novel aspect of ocean boundary layer physics relative to
the atmosphere. Namely, for the atmosphere, radiative absorption is
far less relevant than in the upper ocean, since the atmosphere is
largely transparent to radiation. We therefore consider penetrative
shortwave radiation as distinct from other buoyancy fluxes when
formulating how boundary fluxes impact the ocean.
\subsection{General features of buoyancy forcing}
The buoyancy of a fluid is commonly defined as (e.g., page 83 of
\cite{LargeKPP_lectures})
\begin{equation}
B = -g \, \left( \frac{ \rho - \rho_{o}}{\rho_{o}} \right),
\label{eq:buoyancy-kpp}
\end{equation}
where $g$ is the constant gravitational acceleration, and $\rho_{o}$
is a reference density, taken here to equal the Boussinesq reference
density. A reduction in density is associated with an increase in
buoyancy; that is, the water becomes more {\it buoyant}. Changes in
buoyancy arise through changes in density associated with temperature
and salinity changes, since buoyancy changes are computed relative to
a fixed pressure level. In this way, buoyancy changes are directly
related to processes that impact locally referenced potential density.
Ocean buoyancy is affected through surface ocean heat, salt, and water
fluxes. These fluxes are associated with the following physical
processes.
\begin{itemize}
\item Turbulent processes transfer heat through latent and sensible
heating.
\item Longwave radiation cools the upper ocean, with this radiation
affected by the upper ocean skin temperature.
\item Penetrative shortwave radiation is absorbed in seawater and so
increases buoyancy.
\item The transfer of salt occurs when sea ice melts and forms. This
transfer is proportional to the water mass flux and the difference
in salinity between the liquid ocean and sea ice. More generally,
we simply consider this process to be associated with a salt flux
between sea ice and ocean, with this flux operationally computed as
part of a sea ice model.
\item Advective processes transfer heat and salt across the ocean
surface through the transfer of water mass across the interface.
\end{itemize}
We further detail these processes in the following.
\subsection{Scalar budgets for a surface ocean model grid cell}
Buoyancy is not a prognostic variable in ocean models. So to develop
a quantative understanding of how buoyancy is impacted by surface
fluxes, we consider the evolution of temperature, salinity, and mass
in an arbitrary top model grid cell, and focus exclusively on
evolution arising from surface boundary fluxes. We write these
budgets in their finite volume sense, which includes density and
thickness weighting of scalar tracer fields
\begin{align}
\frac{\partial \, (\rho \, \mathrm{d}z \, \Theta)}{\partial t} &= \Qm \, \Thetam
+ Q_{\theta}^{\mbox{\tiny non-pen}}
+ \left( Q_{\theta}^{\mbox{\tiny pen}}(z=\eta) - Q_{\theta}^{\mbox{\tiny pen}}(z=-\Delta z)
\right)
\label{eq:surface-temperature-equation-kpp}
\\
\frac{\partial \, (\rho \, \mathrm{d}z \, S)}{\partial t} &= \Qm \, \Sm + Q_{S}
\label{eq:surface-salinity-equation-kpp}
\\
\frac{\partial \, (\rho \, \mathrm{d}z)}{\partial t} &= \Qm.
\label{eq:surface-mass-equation-kpp}
\end{align}
We now detail the terms appearing in these equations.
\begin{itemize}
\item $\rho \, \mathrm{d}z$ is the mass per horizontal area of
seawater in the grid cell. For a volume conserving Boussinesq
fluid, the {\it in situ} density, $\rho$, is set to the constant
reference density $\rho_{o}$
\begin{equation}
\rho = \rho_{o} \qquad \mbox{Boussinesq fluid.}
\end{equation}
\item $\Theta$ is the grid cell potential temperature, or more
accurately it is the conservative temperature of
\cite{McDougall2003} (see \cite{TEOS2010}).
\item $S$ is the grid cell salinity.
\item $\Qm$ is the mass flux ($\mbox{kg} \, \mbox{m}^{-2} \,
\mbox{sec}^{-1})$ of water crossing the ocean surface. Following
the sign convention (\ref{eq:boundary-sign-convention}), we
consider $\Qm > 0$ for water entering the ocean (as when
precipitation plus runoff exceeds evaporation).
\item $\Thetam$ is the temperature of water crossing the ocean
surface, and $C_{\mbox{\tiny p}}^{o} \, \Qm \, \Thetam$ is the
associated enthalpy flux ($\mbox{W} \, \mbox{m}^{-2})$. We further
discuss this flux in Section
\ref{subsection:advective-buoyancy-fluxes}.
\item $\Sm$ is the salinity of water crossing the ocean surface, and
$\Qm \, \Sm$ is the associated mass flux of salt. Note that $\Sm$
is typically taken to be zero, as for precipitation and
evaporation. However, rivers can contain a nonzero salt
concentration, so we keep $\Sm$ for the following formulation. We
further discuss this salt flux in Section
\ref{subsection:advective-buoyancy-fluxes}.
\item \label{heat_capacity} $C_{\mbox{\tiny p}}^{o}$ is the seawater heat capacity at
constant pressure ($\mbox{J} \, \mbox{kg}^{-1} \,
\mbox{}^{\circ}\mbox{C}^{-1}$). \cite{TEOS2010} provides the most
precise value appropriate for an ocean with heat measured through
conservative temperature.
\item $Q_{S}$ is the flux of salt ($\mbox{kg} \, \mbox{m}^{-2} \,
\mbox{sec}^{-1})$ that crosses the ocean surface. Following the
sign convention (\ref{eq:boundary-sign-convention}), we take $Q_{S}
> 0$ when salt enters the ocean. This flux arises in the transfer
of salt when sea ice forms and melts. We further discuss this salt
flux in Section \ref{subsection:sea-ice-buoyancy-fluxes}.
\item $C_{\mbox{\tiny p}}^{o} \, Q_{\theta}^{\mbox{\tiny non-pen}}$ is
the non-penetrative surface heat flux associated with turbulent
processes (latent and sensible) and radiative longwave cooling
($\mbox{W} \, \mbox{m}^{-2}$). Following the sign convention
(\ref{eq:boundary-sign-convention}), we take
$Q_{\theta}^{\mbox{\tiny non-pen}} > 0$ for heat entering the ocean
surface (i.e., ocean warming). We further discuss this heat flux in
Section \ref{subsection:non-pen-buoyancy-fluxes}.
\item $C_{\mbox{\tiny p}}^{o} \, Q_{\theta}^{\mbox{\tiny pen}}(z=\eta)$ is the
radiative shortwave heat flux ($\mbox{W} \, \mbox{m}^{-2}$) entering
the ocean through its surface at $z=\eta$, with
$Q_{\theta}^{\mbox{\tiny pen}}(\eta) > 0$ warming the ocean surface.
Likewise, $C_{\mbox{\tiny p}}^{o} \, Q_{\theta}^{\mbox{\tiny pen}}(z=-\Delta z)$ is
the radiative shortwave heat flux leaving the top cell through its
bottom face. We further discuss this heat flux in Section
\ref{subsection:pen-buoyancy-fluxes}.
\end{itemize}
\subsection{Salt fluxes from sea ice melt and formation}
\label{subsection:sea-ice-buoyancy-fluxes}
The mass flux of salt $Q_{S}$ ($\mbox{kg} \; \mbox{m}^{-2} \,
\mbox{sec}^{-1})$ is positive for salt entering the ocean. There is
transport of salt across the ocean surface when sea ice forms and
melts, due to the nonzero salt content in sea ice. Otherwise, the
surface salt flux is generally zero for the large scale ocean. For
ocean models, however, the salt flux can be nonzero when formulating
the surface boundary in terms of virtual salt fluxes rather than real
water fluxes \citep{Huang1993,GriffiesPacSchmidtBalaji2001}. It can
also be non-zero when using an ocean-ice model that is not coupled to
an atmosphere or land model, in which case salt restoring is required
to maintain stability of the overturning circulation (see Section 3 of
\cite{CORE_NYF}).
\subsection{Salt and heat fluxes associated with water transport}
\label{subsection:advective-buoyancy-fluxes}
In most cases, salinity in the water fluxed across the ocean surface
is zero, so that $\Sm=0$. However, there are some cases where rivers
have a nonzero salinity so that $\Sm \ne 0$ and the product $\Qm \,
\Sm$ leads to an advective transport of salt across the ocean surface.
Since water transported across the ocean has a nonzero heat content,
this transport in turn affects the net heat content in the upper
ocean. One can either prescribe the temperature of this water,
$\Thetam$, or the product $\Qm \, \Thetam$. Consider the case where
the product is specified for river water entering the ocean, which is
the case with the GFDL land model used in the earth system model of
\cite{Dunne_etal_part1_2012}. In this case, the heat flux with
respect to $0^{\circ}C$ (in units of $\mbox{W}~\mbox{m}^{-2}$) of
liquid river runoff ${\cal H}^{\mbox{\tiny liquid runoff}}$ is given
to the ocean from the land model, so that
\begin{equation}
\Qm \, \Thetam = \frac{ {\cal H}^{\mbox{\tiny liquid runoff}} } {C_{p}^{\mbox{\tiny liquid runoff}}},
\label{eq:river-heating-kpp}
\end{equation}
with $C_{p}^{\mbox{\tiny liquid runoff}}$ the heat capacity of the
water coming in from the river runoff. Likewise, if the heat
associated with frozen runoff (e.g., calving land ice) is provided by
the land model, then we have
\begin{equation}
\Qm \, \Thetam = \frac{{\cal H}^{\mbox{\tiny solid runoff}}}{C_{p}^{\mbox{\tiny solid runoff}}},
\end{equation}
with $C_{p}^{\mbox{\tiny solid runoff}}$ the heat capacity of the
solid runoff. These two heat capacities are typically provided by the
component model (i.e., the land model) used to compute the runoff
fields. Similar considerations hold for transfer of water between sea
ice models and the ocean.
\subsection{Non-penetrative surface heat fluxes}
\label{subsection:non-pen-buoyancy-fluxes}
Following the sign convention (\ref{eq:boundary-sign-convention}), the
heat flux $C_{\mbox{\tiny p}}^{o} \, Q_{\theta}^{\mbox{\tiny
non-pen}}$ ($\mbox{W} \, \mbox{m}^{-2}$) is positive for heat
entering the ocean. This flux is comprised of the following
contributions \citep[see page 34 of][]{Gill1982}
\begin{equation}
C_{\mbox{\tiny p}}^{o} \, Q_{\theta}^{\mbox{\tiny non-pen}}
= Q_{\mbox{\scriptsize long}} + Q_{\mbox{\scriptsize latent}} +
Q_{\mbox{\scriptsize sens}}.
\label{eq:non-penetrative-for-kpp}
\end{equation}
Longwave, latent, and sensible heat fluxes are typically deposited or
withdrawn from the ocean surface layer (Section
\ref{section:m-o-similarity}). In practice, ocean models assume
these fluxes are taken entirely from the surface grid cell.
These fluxes are termed non-penetrative since they are deposited or
withdrawn from the liquid ocean surface layer, with transport then
occurring through ocean advection and mixing. This behaviour
contrasts to that of penetrative shortwave radiation, which is
transferred into the ocean interior as a function of seawater optics,
so it does not depend on ocean transport. We now comment in a bit
more detail on the various non-penetrative fluxes.
\subsubsection{Longwave radiation}
Longwave radiation leaves the ocean in the form of the
$\sigma_{\mbox{\tiny SB}}\, T^{4}$ Stefan-Boltzmann Law, with $T$ the
skin temperature and
\begin{equation}
\sigma_{\mbox{\tiny SB}} = 5.6734 \, \times \, 10^{-8}~\mbox{W}~\mbox{m}^{-2}~\mbox{}^{\circ}\mbox{K}^{-4}
\end{equation}
the Stefan-Boltzmann constant. Following the sign convention detailed
in Section \ref{subsection:conventions}, $Q_{\mbox{\scriptsize long}}
<0$ since the longwave heat flux removes heat from the ocean surface
and sends it back to the atmosphere.
\subsubsection{Latent heat fluxes}
$Q_{\mbox{\scriptsize latent}}$ arises from phase changes whereby
liquid seawater either evaporates, or it acts to melt frozen
precipitation. In either case, $Q_{\mbox{\scriptsize latent}} < 0$
since the liquid ocean loses heat to energize the phase changes.
When seawater evaporates, the latent heat lost by the ocean is
determined by the latent heat of vaporization for fresh water
\begin{equation}
H^{\mbox{\tiny vapor}} = 2.5 \times 10^{6} \, \mbox{J} \; \mbox{kg}^{-1},
\label{eq:latent-heat-vapor}
\end{equation}
so that
\begin{equation}
Q_{\mbox{\tiny evap}} = H^{\mbox{\tiny vapor}} \, \Qm^{\mbox{\tiny evap}}
\end{equation}
where $\Qm^{\mbox{\tiny evap}} < 0$ is the mass flux ($\mbox{kg} \,
\mbox{m}^{-2} \; \mbox{sec}^{-1})$ of fresh water leaving the ocean
due to evaporation. A similar expression holds when seawater melts
frozen precipitation (e.g., snow), in which case
\begin{equation}
H^{\mbox{\tiny fusion}} = 3.34 \times 10^{5} \, \mbox{J} \; \mbox{kg}^{-1},
\label{eq:latent-heat-fusion}
\end{equation}
so that
\begin{equation}
Q_{\mbox{\tiny melt}} = -H^{\mbox{\tiny fusion}} \, \Qm^{\mbox{\tiny frozen precip}},
\end{equation}
where $\Qm^{\mbox{\tiny frozen precip}} > 0$ is the mass flux
($\mbox{kg} \; \mbox{m}^{-2} \, \mbox{sec}^{-1})$ of frozen
precipitation falling onto the ocean surface. Again, both
$Q_{\mbox{\tiny evap}}$ and $Q_{\mbox{\tiny melt}}$ are negative since
latent heating extracts heat from the ocean.
\subsubsection{Sensible heat fluxes}
$Q_{\mbox{\tiny sens}}$ is the sensible heat transfer proportional to
the difference between the ocean temperature and that of the
atmosphere, sea ice, or land ice. Sensible heating generally acts to
cool the ocean ($Q_{\mbox{\tiny sens}} < 0$), particularly near
western boundary currents such as the Gulf Stream, Kuroshio, and