/[MITgcm]/MITgcm_contrib/articles/ceaice/ceaice.tex
ViewVC logotype

Contents of /MITgcm_contrib/articles/ceaice/ceaice.tex

Parent Directory Parent Directory | Revision Log Revision Log | View Revision Graph Revision Graph


Revision 1.15 - (show annotations) (download) (as text)
Tue Feb 26 17:21:48 2008 UTC (17 years, 4 months ago) by mlosch
Branch: MAIN
Changes since 1.14: +48 -32 lines
File MIME type: application/x-tex
fix a few units, spelling etc.
add a paragraph about lateral boundary conditions
add a few comments about advection scheme (and something for JMC to
comment on)

1 % $Header: /u/gcmpack/MITgcm_contrib/articles/ceaice/ceaice.tex,v 1.14 2008/02/26 00:13:20 dimitri Exp $
2 % $Name: $
3 \documentclass[12pt]{article}
4
5 \usepackage[]{graphicx}
6 \usepackage{subfigure}
7
8 \usepackage[round,comma]{natbib}
9 \bibliographystyle{bib/agu04}
10
11 \usepackage{amsmath,amssymb}
12 \newcommand\bmmax{10} \newcommand\hmmax{10}
13 \usepackage{bm}
14
15 % math abbreviations
16 \newcommand{\vek}[1]{\ensuremath{\mathbf{#1}}}
17 \newcommand{\mat}[1]{\ensuremath{\mathbf{#1}}}
18 \newcommand{\vtau}{\bm{{\tau}}}
19
20 \newcommand{\degree}{\ensuremath{^\circ}}
21 \newcommand{\degC}{\,\ensuremath{\degree}C}
22 \newcommand{\degE}{\ensuremath{\degree}\,E}
23 \newcommand{\degS}{\ensuremath{\degree}\,S}
24 \newcommand{\degN}{\ensuremath{\degree}\,N}
25 \newcommand{\degW}{\ensuremath{\degree}\,W}
26
27 % cross reference scheme
28 \newcommand{\reffig}[1]{Figure~\ref{fig:#1}}
29 \newcommand{\reftab}[1]{Table~\ref{tab:#1}}
30 \newcommand{\refapp}[1]{Appendix~\ref{app:#1}}
31 \newcommand{\refsec}[1]{Section~\ref{sec:#1}}
32 \newcommand{\refeq}[1]{\,(\ref{eq:#1})}
33 \newcommand{\refeqs}[2]{\,(\ref{eq:#1})--(\ref{eq:#2})}
34
35 \newlength{\stdfigwidth}\setlength{\stdfigwidth}{20pc}
36 %\newlength{\stdfigwidth}\setlength{\stdfigwidth}{\columnwidth}
37 \newlength{\mediumfigwidth}\setlength{\mediumfigwidth}{39pc}
38 %\newlength{\widefigwidth}\setlength{\widefigwidth}{39pc}
39 \newlength{\widefigwidth}\setlength{\widefigwidth}{\textwidth}
40 \newcommand{\fpath}{figs}
41
42 % commenting scheme
43 \newcommand{\ml}[1]{\textsf{\slshape #1}}
44
45 \title{A Dynamic-Thermodynamic Sea ice Model for Ocean Climate
46 Estimation on an Arakawa C-Grid}
47
48 \author{Martin Losch, Dimitris Menemenlis, Patrick Heimbach, \\
49 Jean-Michel Campin, and Chris Hill}
50 \begin{document}
51
52 \maketitle
53
54 \begin{abstract}
55 As part of ongoing efforts to obtain a best possible synthesis of most
56 available, global-scale, ocean and sea ice data, a dynamic and thermodynamic
57 sea-ice model has been coupled to the Massachusetts Institute of Technology
58 general circulation model (MITgcm). Ice mechanics follow a viscous plastic
59 rheology and the ice momentum equations are solved numerically using either
60 line successive relaxation (LSR) or elastic-viscous-plastic (EVP) dynamic
61 models. Ice thermodynamics are represented using either a zero-heat-capacity
62 formulation or a two-layer formulation that conserves enthalpy. The model
63 includes prognostic variables for snow and for sea-ice salinity. The above
64 sea ice model components were borrowed from current-generation climate models
65 but they were reformulated on an Arakawa C-grid in order to match the MITgcm
66 oceanic grid and they were modified in many ways to permit efficient and
67 accurate automatic differentiation. This paper describes the MITgcm sea ice
68 model; it presents example Arctic and Antarctic results from a realistic,
69 eddy-permitting, global ocean and sea-ice configuration; it compares B-grid
70 and C-grid dynamic solvers in a regional Arctic configuration; and it presents
71 example results from coupled ocean and sea-ice adjoint-model integrations.
72
73 \end{abstract}
74
75 \section{Introduction}
76 \label{sec:intro}
77
78 The availability of an adjoint model as a powerful research
79 tool complementary to an ocean model was a major design
80 requirement early on in the development of the MIT general
81 circulation model (MITgcm) [Marshall et al. 1997a,
82 Marotzke et al. 1999, Adcroft et al. 2002]. It was recognized
83 that the adjoint permitted very efficient computation
84 of gradients of various scalar-valued model diagnostics,
85 norms or, generally, objective functions with respect
86 to external or independent parameters. Such gradients
87 arise in at least two major contexts. If the objective function
88 is the sum of squared model vs. obervation differences
89 weighted by e.g. the inverse error covariances, the gradient
90 of the objective function can be used to optimize this measure
91 of model vs. data misfit in a least-squares sense. One
92 is then solving a problem of statistical state estimation.
93 If the objective function is a key oceanographic quantity
94 such as meridional heat or volume transport, ocean heat
95 content or mean surface temperature index, the gradient
96 provides a complete set of sensitivities of this quantity
97 with respect to all independent variables simultaneously.
98
99 References to existing sea-ice adjoint models, explaining that they are either
100 for simplified configurations, for ice-only studies, or for short-duration
101 studies to motivate the present work.
102
103 Traditionally, probably for historical reasons and the ease of
104 treating the Coriolis term, most standard sea-ice models are
105 discretized on Arakawa-B-grids \citep[e.g.,][]{hibler79, harder99,
106 kreyscher00, zhang98, hunke97}. From the perspective of coupling a
107 sea ice-model to a C-grid ocean model, the exchange of fluxes of heat
108 and fresh-water pose no difficulty for a B-grid sea-ice model
109 \citep[e.g.,][]{timmermann02a}. However, surface stress is defined at
110 velocities points and thus needs to be interpolated between a B-grid
111 sea-ice model and a C-grid ocean model. While the smoothing implicitly
112 associated with this interpolation may mask grid scale noise, it may
113 in two-way coupling lead to a computational mode as will be shown. By
114 choosing a C-grid for the sea-ice model, we circumvent this difficulty
115 altogether and render the stress coupling as consistent as the
116 buoyancy coupling.
117
118 A further advantage of the C-grid formulation is apparent in narrow
119 straits. In the limit of only one grid cell between coasts there is no
120 flux allowed for a B-grid (with no-slip lateral boundary counditions),
121 whereas the C-grid formulation allows a flux of sea-ice through this
122 passage for all types of lateral boundary conditions. We
123 demonstrate this effect in the Candian archipelago.
124
125 Talk about problems that make the sea-ice-ocean code very sensitive and
126 changes in the code that reduce these sensitivities.
127
128 This paper describes the MITgcm sea ice
129 model; it presents example Arctic and Antarctic results from a realistic,
130 eddy-permitting, global ocean and sea-ice configuration; it compares B-grid
131 and C-grid dynamic solvers in a regional Arctic configuration; and it presents
132 example results from coupled ocean and sea-ice adjoint-model integrations.
133
134 \section{Model}
135 \label{sec:model}
136
137 \subsection{Dynamics}
138 \label{sec:dynamics}
139
140 The momentum equation of the sea-ice model is
141 \begin{equation}
142 \label{eq:momseaice}
143 m \frac{D\vek{u}}{Dt} = -mf\vek{k}\times\vek{u} + \vtau_{air} +
144 \vtau_{ocean} - m \nabla{\phi(0)} + \vek{F},
145 \end{equation}
146 where $m=m_{i}+m_{s}$ is the ice and snow mass per unit area;
147 $\vek{u}=u\vek{i}+v\vek{j}$ is the ice velocity vector;
148 $\vek{i}$, $\vek{j}$, and $\vek{k}$ are unit vectors in the $x$, $y$, and $z$
149 directions, respectively;
150 $f$ is the Coriolis parameter;
151 $\vtau_{air}$ and $\vtau_{ocean}$ are the wind-ice and ocean-ice stresses,
152 respectively;
153 $g$ is the gravity accelation;
154 $\nabla\phi(0)$ is the gradient (or tilt) of the sea surface height;
155 $\phi(0) = g\eta + p_{a}/\rho_{0}$ is the sea surface height potential
156 in response to ocean dynamics ($g\eta$) and to atmospheric pressure
157 loading ($p_{a}/\rho_{0}$, where $\rho_{0}$ is a reference density);
158 and $\vek{F}=\nabla\cdot\sigma$ is the divergence of the internal ice stress
159 tensor $\sigma_{ij}$.
160 When using the rescaled vertical coordinate system, z$^\ast$, of
161 \citet{cam08}, $\phi(0)$ also includes a term due to snow and ice
162 loading, $mg/\rho_{0}$.
163 Advection of sea-ice momentum is neglected. The wind and ice-ocean stress
164 terms are given by
165 \begin{align*}
166 \vtau_{air} = & \rho_{air} C_{air} |\vek{U}_{air} -\vek{u}|
167 R_{air} (\vek{U}_{air} -\vek{u}), \\
168 \vtau_{ocean} = & \rho_{ocean}C_{ocean} |\vek{U}_{ocean}-\vek{u}|
169 R_{ocean}(\vek{U}_{ocean}-\vek{u}), \\
170 \end{align*}
171 where $\vek{U}_{air/ocean}$ are the surface winds of the atmosphere
172 and surface currents of the ocean, respectively; $C_{air/ocean}$ are
173 air and ocean drag coefficients; $\rho_{air/ocean}$ are reference
174 densities; and $R_{air/ocean}$ are rotation matrices that act on the
175 wind/current vectors.
176
177 For an isotropic system the stress tensor $\sigma_{ij}$ ($i,j=1,2$) can
178 be related to the ice strain rate and strength by a nonlinear
179 viscous-plastic (VP) constitutive law \citep{hibler79, zhang98}:
180 \begin{equation}
181 \label{eq:vpequation}
182 \sigma_{ij}=2\eta(\dot{\epsilon}_{ij},P)\dot{\epsilon}_{ij}
183 + \left[\zeta(\dot{\epsilon}_{ij},P) -
184 \eta(\dot{\epsilon}_{ij},P)\right]\dot{\epsilon}_{kk}\delta_{ij}
185 - \frac{P}{2}\delta_{ij}.
186 \end{equation}
187 The ice strain rate is given by
188 \begin{equation*}
189 \dot{\epsilon}_{ij} = \frac{1}{2}\left(
190 \frac{\partial{u_{i}}}{\partial{x_{j}}} +
191 \frac{\partial{u_{j}}}{\partial{x_{i}}}\right).
192 \end{equation*}
193 The maximum ice pressure $P_{\max}$, a measure of ice strength, depends on
194 both thickness $h$ and compactness (concentration) $c$:
195 \begin{equation}
196 P_{\max} = P^{*}c\,h\,e^{[C^{*}\cdot(1-c)]},
197 \label{eq:icestrength}
198 \end{equation}
199 with the constants $P^{*}$ and $C^{*}$. The nonlinear bulk and shear
200 viscosities $\eta$ and $\zeta$ are functions of ice strain rate
201 invariants and ice strength such that the principal components of the
202 stress lie on an elliptical yield curve with the ratio of major to
203 minor axis $e$ equal to $2$; they are given by:
204 \begin{align*}
205 \zeta =& \min\left(\frac{P_{\max}}{2\max(\Delta,\Delta_{\min})},
206 \zeta_{\max}\right) \\
207 \eta =& \frac{\zeta}{e^2} \\
208 \intertext{with the abbreviation}
209 \Delta = & \left[
210 \left(\dot{\epsilon}_{11}^2+\dot{\epsilon}_{22}^2\right)
211 (1+e^{-2}) + 4e^{-2}\dot{\epsilon}_{12}^2 +
212 2\dot{\epsilon}_{11}\dot{\epsilon}_{22} (1-e^{-2})
213 \right]^{-\frac{1}{2}}
214 \end{align*}
215 The bulk viscosities are bounded above by imposing both a minimum
216 $\Delta_{\min}=10^{-11}\text{\,s}^{-1}$ (for numerical reasons) and a
217 maximum $\zeta_{\max} = P_{\max}/\Delta^*$, where
218 $\Delta^*=(5\times10^{12}/2\times10^4)\text{\,s}^{-1}$. For stress
219 tensor computation the replacement pressure $P = 2\,\Delta\zeta$
220 \citep{hibler95} is used so that the stress state always lies on the
221 elliptic yield curve by definition.
222
223 In the so-called truncated ellipse method the shear viscosity $\eta$
224 is capped to suppress any tensile stress \citep{hibler97, geiger98}:
225 \begin{equation}
226 \label{eq:etatem}
227 \eta = \min\left(\frac{\zeta}{e^2},
228 \frac{\frac{P}{2}-\zeta(\dot{\epsilon}_{11}+\dot{\epsilon}_{22})}
229 {\sqrt{(\dot{\epsilon}_{11}+\dot{\epsilon}_{22})^2
230 +4\dot{\epsilon}_{12}^2}}\right).
231 \end{equation}
232
233 In the current implementation, the VP-model is integrated with the
234 semi-implicit line successive over relaxation (LSOR)-solver of
235 \citet{zhang98}, which allows for long time steps that, in our case,
236 are limited by the explicit treatment of the Coriolis term. The
237 explicit treatment of the Coriolis term does not represent a severe
238 limitation because it restricts the time step to approximately the
239 same length as in the ocean model where the Coriolis term is also
240 treated explicitly.
241
242 \citet{hunke97}'s introduced an elastic contribution to the strain
243 rate in order to regularize Eq.\refeq{vpequation} in such a way that
244 the resulting elastic-viscous-plastic (EVP) and VP models are
245 identical at steady state,
246 \begin{equation}
247 \label{eq:evpequation}
248 \frac{1}{E}\frac{\partial\sigma_{ij}}{\partial{t}} +
249 \frac{1}{2\eta}\sigma_{ij}
250 + \frac{\eta - \zeta}{4\zeta\eta}\sigma_{kk}\delta_{ij}
251 + \frac{P}{4\zeta}\delta_{ij}
252 = \dot{\epsilon}_{ij}.
253 \end{equation}
254 %In the EVP model, equations for the components of the stress tensor
255 %$\sigma_{ij}$ are solved explicitly. Both model formulations will be
256 %used and compared the present sea-ice model study.
257 The EVP-model uses an explicit time stepping scheme with a short
258 timestep. According to the recommendation of \citet{hunke97}, the
259 EVP-model is stepped forward in time 120 times within the physical
260 ocean model time step (although this parameter is under debate), to
261 allow for elastic waves to disappear. Because the scheme does not
262 require a matrix inversion it is fast in spite of the small timestep
263 \citep{hunke97}. For completeness, we repeat the equations for the
264 components of the stress tensor $\sigma_{1} =
265 \sigma_{11}+\sigma_{22}$, $\sigma_{2}= \sigma_{11}-\sigma_{22}$, and
266 $\sigma_{12}$. Introducing the divergence $D_D =
267 \dot{\epsilon}_{11}+\dot{\epsilon}_{22}$, and the horizontal tension
268 and shearing strain rates, $D_T =
269 \dot{\epsilon}_{11}-\dot{\epsilon}_{22}$ and $D_S =
270 2\dot{\epsilon}_{12}$, respectively, and using the above abbreviations,
271 the equations can be written as:
272 \begin{align}
273 \label{eq:evpstresstensor1}
274 \frac{\partial\sigma_{1}}{\partial{t}} + \frac{\sigma_{1}}{2T} +
275 \frac{P}{2T} &= \frac{P}{2T\Delta} D_D \\
276 \label{eq:evpstresstensor2}
277 \frac{\partial\sigma_{2}}{\partial{t}} + \frac{\sigma_{2} e^{2}}{2T}
278 &= \frac{P}{2T\Delta} D_T \\
279 \label{eq:evpstresstensor12}
280 \frac{\partial\sigma_{12}}{\partial{t}} + \frac{\sigma_{12} e^{2}}{2T}
281 &= \frac{P}{4T\Delta} D_S
282 \end{align}
283 Here, the elastic parameter $E$ is redefined in terms of a damping timescale
284 $T$ for elastic waves \[E=\frac{\zeta}{T}.\]
285 $T=E_{0}\Delta{t}$ with the tunable parameter $E_0<1$ and
286 the external (long) timestep $\Delta{t}$. \citet{hunke97} recommend
287 $E_{0} = \frac{1}{3}$.
288
289 For details of the spatial discretization, the reader is referred to
290 \citet{zhang98, zhang03}. Our discretization differs only (but
291 importantly) in the underlying grid, namely the Arakawa C-grid, but is
292 otherwise straightforward. The EVP model, in particular, is discretized
293 naturally on the C-grid with $\sigma_{1}$ and $\sigma_{2}$ on the
294 center points and $\sigma_{12}$ on the corner (or vorticity) points of
295 the grid. With this choice all derivatives are discretized as central
296 differences and averaging is only involved in computing $\Delta$ and
297 $P$ at vorticity points.
298
299 For a general curvilinear grid, one needs in principle to take metric
300 terms into account that arise in the transformation of a curvilinear
301 grid on the sphere. For now, however, we can neglect these metric
302 terms because they are very small on the \ml{[modify following
303 section3:] cubed sphere grids used in this paper; in particular,
304 only near the edges of the cubed sphere grid, we expect them to be
305 non-zero, but these edges are at approximately 35\degS\ or 35\degN\
306 which are never covered by sea-ice in our simulations. Everywhere
307 else the coordinate system is locally nearly cartesian.} However, for
308 last-glacial-maximum or snowball-earth-like simulations the question
309 of metric terms needs to be reconsidered. Either, one includes these
310 terms as in \citet{zhang03}, or one finds a vector-invariant
311 formulation for the sea-ice internal stress term that does not require
312 any metric terms, as it is done in the ocean dynamics of the MITgcm
313 \citep{adcroft04:_cubed_sphere}.
314
315 Lateral boundary conditions are naturally ``no-slip'' for B-grids, as
316 the tangential velocities points lie on the boundary. For C-grids, the
317 lateral boundary condition for tangential velocities is realized via
318 ``ghost points'', allowing alternatively no-slip or free-slip
319 conditions. In ocean models free-slip boundary conditions in
320 conjunction with piecewise-constant (``castellated'') coastlines have
321 been shown to reduce in effect to no-slip boundary conditions
322 \citep{adcroft98:_slippery_coast}; for sea-ice models the effects of
323 lateral boundary conditions have not yet been studied.
324
325 Moving sea ice exerts a stress on the ocean which is the opposite of
326 the stress $\vtau_{ocean}$ in Eq.\refeq{momseaice}. This stess is
327 applied directly to the surface layer of the ocean model. An
328 alternative ocean stress formulation is given by \citet{hibler87}.
329 Rather than applying $\vtau_{ocean}$ directly, the stress is derived
330 from integrating over the ice thickness to the bottom of the oceanic
331 surface layer. In the resulting equation for the \emph{combined}
332 ocean-ice momentum, the interfacial stress cancels and the total
333 stress appears as the sum of windstress and divergence of internal ice
334 stresses: $\delta(z) (\vtau_{air} + \vek{F})/\rho_0$, \citep[see also
335 Eq.\,2 of][]{hibler87}. The disadvantage of this formulation is that
336 now the velocity in the surface layer of the ocean that is used to
337 advect tracers, is really an average over the ocean surface
338 velocity and the ice velocity leading to an inconsistency as the ice
339 temperature and salinity are different from the oceanic variables.
340
341 Sea ice distributions are characterized by sharp gradients and edges.
342 For this reason, we employ positive, multidimensional 2nd and 3rd-order
343 advection scheme with flux limiter \citep{roe85, hundsdorfer94} for the
344 thermodynamic variables discussed in the next section.
345
346 \subparagraph{boundary conditions: no-slip, free-slip, half-slip}
347
348 \begin{itemize}
349 \item transition from existing B-Grid to C-Grid
350 \item boundary conditions: no-slip, free-slip, half-slip
351 \item fancy (multi dimensional) advection schemes
352 \item VP vs.\ EVP \citep{hunke97}
353 \item ocean stress formulation \citep{hibler87}
354 \end{itemize}
355
356 \subsection{Thermodynamics}
357 \label{sec:thermodynamics}
358
359 In the original formulation the sea ice model \citep{menemenlis05}
360 uses simple thermodynamics following the appendix of
361 \citet{semtner76}. This formulation does not allow storage of heat
362 (heat capacity of ice is zero, and this type of model is often refered
363 to as a ``zero-layer'' model). Upward heat flux is parameterized
364 assuming a linear temperature profile and together with a constant ice
365 conductivity. It is expressed as $(K/h)(T_{w}-T_{0})$, where $K$ is
366 the ice conductivity, $h$ the ice thickness, and $T_{w}-T_{0}$ the
367 difference between water and ice surface temperatures. The surface
368 heat budget is computed in a similar way to that of
369 \citet{parkinson79} and \citet{manabe79}.
370
371 There is considerable doubt about the reliability of such a simple
372 thermodynamic model---\citet{semtner84} found significant errors in
373 phase (one month lead) and amplitude ($\approx$50\%\,overestimate) in
374 such models---, so that today many sea ice models employ more complex
375 thermodynamics. A popular thermodynamics model of \citet{winton00} is
376 based on the 3-layer model of \citet{semtner76} and treats brine
377 content by means of enthalphy conservation. This model requires in
378 addition to ice-thickness and compactness (fractional area) additional
379 state variables to be advected by ice velocities, namely enthalphy of
380 the two ice layers and the thickness of the overlying snow layer.
381 \ml{[Jean-Michel, your turf: ]Care must be taken in advecting these
382 quantities in order to ensure conservation of enthalphy. Currently
383 this can only be accomplished with a 2nd-order advection scheme with
384 flux limiter \citep{roe85}.}
385
386
387 \subsection{C-grid}
388 \begin{itemize}
389 \item no-slip vs. free-slip for both lsr and evp;
390 "diagnostics" to look at and use for comparison
391 \begin{itemize}
392 \item ice transport through Fram Strait/Denmark Strait/Davis
393 Strait/Bering strait (these are general)
394 \item ice transport through narrow passages, e.g.\ Nares-Strait
395 \end{itemize}
396 \item compare different advection schemes (if lsr turns out to be more
397 effective, then with lsr otherwise I prefer evp), eg.
398 \begin{itemize}
399 \item default 2nd-order with diff1=0.002
400 \item 1st-order upwind with diff1=0.
401 \item DST3FL (SEAICEadvScheme=33 with diff1=0., should work, works for me)
402 \item 2nd-order wit flux limiter (SEAICEadvScheme=77 with diff1=0.)
403 \end{itemize}
404 That should be enough. Here, total ice mass and location of ice edge
405 is interesting. However, this comparison can be done in an idealized
406 domain, may not require full Arctic Domain?
407 \item
408 Do a little study on the parameters of LSR and EVP
409 \begin{enumerate}
410 \item convergence of LSR, how many iterations do you need to get a
411 true elliptic yield curve
412 \item vary deltaTevp and the relaxation parameter for EVP and see when
413 the EVP solution breaks down (relative to the forcing time scale).
414 For this, it is essential that the evp solver gives use "stripeless"
415 solutions, that is your dtevp = 1sec solutions/or 10sec solutions
416 with SEAICE\_evpDampC = 615.
417 \end{enumerate}
418 \end{itemize}
419
420 \section{Forward sensitivity experiments}
421 \label{sec:forward}
422
423 A second series of forward sensitivity experiments have been carried out on an
424 Arctic Ocean domain with open boundaries. Once again the objective is to
425 compare the old B-grid LSR dynamic solver with the new C-grid LSR and EVP
426 solvers. One additional experiment is carried out to illustrate the
427 differences between the two main options for sea ice thermodynamics in the MITgcm.
428
429 \subsection{Arctic Domain with Open Boundaries}
430 \label{sec:arctic}
431
432 The Arctic domain of integration is illustrated in Fig.~\ref{fig:arctic1}. It
433 is carved out from, and obtains open boundary conditions from, the
434 global cubed-sphere configuration of the Estimating the Circulation
435 and Climate of the Ocean, Phase II (ECCO2) project
436 \citet{menemenlis05}. The domain size is 420 by 384 grid boxes
437 horizontally with mean horizontal grid spacing of 18 km.
438
439 \begin{figure}
440 %\centerline{{\includegraphics*[width=0.44\linewidth]{\fpath/arctic1.eps}}}
441 \caption{Bathymetry of Arctic Domain.\label{fig:arctic1}}
442 \end{figure}
443
444 There are 50 vertical levels ranging in thickness from 10 m near the surface
445 to approximately 450 m at a maximum model depth of 6150 m. Bathymetry is from
446 the National Geophysical Data Center (NGDC) 2-minute gridded global relief
447 data (ETOPO2) and the model employs the partial-cell formulation of
448 \citet{adcroft97:_shaved_cells}, which permits accurate representation of the bathymetry. The
449 model is integrated in a volume-conserving configuration using a finite volume
450 discretization with C-grid staggering of the prognostic variables. In the
451 ocean, the non-linear equation of state of \citet{jackett95}. The ocean model is
452 coupled to a sea-ice model described hereinabove.
453
454 This particular ECCO2 simulation is initialized from rest using the
455 January temperature and salinity distribution from the World Ocean
456 Atlas 2001 (WOA01) [Conkright et al., 2002] and it is integrated for
457 32 years prior to the 1996--2001 period discussed in the study. Surface
458 boundary conditions are from the National Centers for Environmental
459 Prediction and the National Center for Atmospheric Research
460 (NCEP/NCAR) atmospheric reanalysis [Kistler et al., 2001]. Six-hourly
461 surface winds, temperature, humidity, downward short- and long-wave
462 radiations, and precipitation are converted to heat, freshwater, and
463 wind stress fluxes using the \citet{large81, large82} bulk formulae.
464 Shortwave radiation decays exponentially as per Paulson and Simpson
465 [1977]. Additionally the time-mean river run-off from Large and Nurser
466 [2001] is applied and there is a relaxation to the monthly-mean
467 climatological sea surface salinity values from WOA01 with a
468 relaxation time scale of 3 months. Vertical mixing follows
469 \citet{large94} with background vertical diffusivity of
470 $1.5\times10^{-5}\text{\,m$^{2}$\,s$^{-1}$}$ and viscosity of
471 $10^{-3}\text{\,m$^{2}$\,s$^{-1}$}$. A third order, direct-space-time
472 advection scheme with flux limiter is employed \citep{hundsdorfer94}
473 and there is no explicit horizontal diffusivity. Horizontal viscosity
474 follows \citet{lei96} but
475 modified to sense the divergent flow as per Fox-Kemper and Menemenlis
476 [in press]. Shortwave radiation decays exponentially as per Paulson
477 and Simpson [1977]. Additionally, the time-mean runoff of Large and
478 Nurser [2001] is applied near the coastline and, where there is open
479 water, there is a relaxation to monthly-mean WOA01 sea surface
480 salinity with a time constant of 45 days.
481
482 Open water, dry
483 ice, wet ice, dry snow, and wet snow albedo are, respectively, 0.15, 0.85,
484 0.76, 0.94, and 0.8.
485
486 \begin{itemize}
487 \item Configuration
488 \item OBCS from cube
489 \item forcing
490 \item 1/2 and full resolution
491 \item with a few JFM figs from C-grid LSR no slip
492 ice transport through Canadian Archipelago
493 thickness distribution
494 ice velocity and transport
495 \end{itemize}
496
497 \begin{itemize}
498 \item Arctic configuration
499 \item ice transport through straits and near boundaries
500 \item focus on narrow straits in the Canadian Archipelago
501 \end{itemize}
502
503 \begin{itemize}
504 \item B-grid LSR no-slip
505 \item C-grid LSR no-slip
506 \item C-grid LSR slip
507 \item C-grid EVP no-slip
508 \item C-grid EVP slip
509 \item C-grid LSR + TEM (truncated ellipse method, no tensile stress, new flag)
510 \item C-grid LSR no-slip + Winton
511 \item speed-performance-accuracy (small)
512 ice transport through Canadian Archipelago differences
513 thickness distribution differences
514 ice velocity and transport differences
515 \end{itemize}
516
517 We anticipate small differences between the different models due to:
518 \begin{itemize}
519 \item advection schemes: along the ice-edge and regions with large
520 gradients
521 \item C-grid: less transport through narrow straits for no slip
522 conditons, more for free slip
523 \item VP vs.\ EVP: speed performance, accuracy?
524 \item ocean stress: different water mass properties beneath the ice
525 \end{itemize}
526
527 \section{Adjoint sensiivities of the MITsim}
528
529 \subsection{The adjoint of MITsim}
530
531 The ability to generate tangent linear and adjoint model components
532 of the MITsim has been a main design task.
533 For the ocean the adjoint capability has proven to be an
534 invaluable tool for sensitivity analysis as well as state estimation.
535 In short, the adjoint enables very efficient computation of the gradient
536 of scalar-valued model diagnostics (called cost function or objective function)
537 with respect to many model "variables".
538 These variables can be two- or three-dimensional fields of initial
539 conditions, model parameters such as mixing coefficients, or
540 time-varying surface or lateral (open) boundary conditions.
541 When combined, these variables span a potentially high-dimensional
542 (e.g. O(10$^8$)) so-called control space. Performing parameter perturbations
543 to assess model sensitivities quickly becomes prohibitive at these scales.
544 Alternatively, (time-varying) sensitivities of the objective function
545 to any element of the control space can be computed very efficiently in
546 one single adjoint
547 model integration, provided an efficient adjoint model is available.
548
549 [REFERENCES]
550
551
552 The adjoint operator (ADM) is the transpose of the tangent linear operator (TLM)
553 of the full (in general nonlinear) forward model, i.e. the MITsim.
554 The TLM maps perturbations of elements of the control space
555 (e.g. initial ice thickness distribution)
556 via the model Jacobian
557 to a perturbation in the objective function
558 (e.g. sea-ice export at the end of the integration interval).
559 \textit{Tangent} linearity ensures that the derivatives are evaluated
560 with respect to the underlying model trajectory at each point in time.
561 This is crucial for nonlinear trajectories and the presence of different
562 regimes (e.g. effect of the seaice growth term at or away from the
563 freezing point of the ocean surface).
564 Ensuring tangent linearity can be easily achieved by integrating
565 the full model in sync with the TLM to provide the underlying model state.
566 Ensuring \textit{tangent} adjoints is equally crucial, but much more
567 difficult to achieve because of the reverse nature of the integration:
568 the adjoint accumulates sensitivities backward in time,
569 starting from a unit perturbation of the objective function.
570 The adjoint model requires the model state in reverse order.
571 This presents one of the major complications in deriving an
572 exact, i.e. \textit{tangent} adjoint model.
573
574 Following closely the development and maintenance of TLM and ADM
575 components of the MITgcm we have relied heavily on the
576 autmomatic differentiation (AD) tool
577 "Transformation of Algorithms in Fortran" (TAF)
578 developed by Fastopt (Giering and Kaminski, 1998)
579 to derive TLM and ADM code of the MITsim.
580 Briefly, the nonlinear parent model is fed to the AD tool which produces
581 derivative code for the specified control space and objective function.
582 Following this approach has (apart from its evident success)
583 several advantages:
584 (1) the adjoint model is the exact adjoint operator of the parent model,
585 (2) the adjoint model can be kept up to date with respect to ongoing
586 development of the parent model, and adjustments to the parent model
587 to extend the automatically generated adjoint are incremental changes
588 only, rather than extensive re-developments,
589 (3) the parallel structure of the parent model is preserved
590 by the adjoint model, ensuring efficient use in high performance
591 computing environments.
592
593 Some initial code adjustments are required to support dependency analysis
594 of the flow reversal and certain language limitations which may lead
595 to irreducible flow graphs (e.g. GOTO statements).
596 The problem of providing the required model state in reverse order
597 at the time of evaluating nonlinear or conditional
598 derivatives is solved via balancing
599 storing vs. recomputation of the model state in a multi-level
600 checkpointing loop.
601 Again, an initial code adjustment is required to support TAFs
602 checkpointing capability.
603 The code adjustments are sufficiently simple so as not to cause
604 major limitations to the full nonlinear parent model.
605 Once in place, an adjoint model of a new model configuration
606 may be derived in about 10 minutes.
607
608 [HIGHLIGHT COUPLED NATURE OF THE ADJOINT!]
609
610 \subsection{Special considerations}
611
612 * growth term(?)
613
614 * small active denominators
615
616 * dynamic solver (implicit function theorem)
617
618 * approximate adjoints
619
620
621 \subsection{An example: sensitivities of sea-ice export through Fram Strait}
622
623 We demonstrate the power of the adjoint method
624 in the context of investigating sea-ice export sensitivities through Fram Strait
625 (for details of this study see Heimbach et al., 2007).
626 %\citep[for details of this study see][]{heimbach07}. %Heimbach et al., 2007).
627 The domain chosen is a coarsened version of the Arctic face of the
628 high-resolution cubed-sphere configuration of the ECCO2 project
629 \citep[see][]{menemenlis05}. It covers the entire Arctic,
630 extends into the North Pacific such as to cover the entire
631 ice-covered regions, and comprises parts of the North Atlantic
632 down to XXN to enable analysis of remote influences of the
633 North Atlantic current to sea-ice variability and export.
634 The horizontal resolution varies between XX and YY km
635 with 50 unevenly spaced vertical levels.
636 The adjoint models run efficiently on 80 processors
637 (benchmarks have been performed both on an SGI Altix as well as an
638 IBM SP5 at NASA/ARC).
639
640 Following a 1-year spinup, the model has been integrated for four
641 years between 1992 and 1995. It is forced using realistic 6-hourly
642 NCEP/NCAR atmospheric state variables. Over the open ocean these are
643 converted into air-sea fluxes via the bulk formulae of
644 \citet{large04}. Derivation of air-sea fluxes in the presence of
645 sea-ice is handled by the ice model as described in \refsec{model}.
646 The objective function chosen is sea-ice export through Fram Strait
647 computed for December 1995. The adjoint model computes sensitivities
648 to sea-ice export back in time from 1995 to 1992 along this
649 trajectory. In principle all adjoint model variable (i.e., Lagrange
650 multipliers) of the coupled ocean/sea-ice model are available to
651 analyze the transient sensitivity behaviour of the ocean and sea-ice
652 state. Over the open ocean, the adjoint of the bulk formula scheme
653 computes sensitivities to the time-varying atmospheric state. Over
654 ice-covered parts, the sea-ice adjoint converts surface ocean
655 sensitivities to atmospheric sensitivities.
656
657 \reffig{4yradjheff}(a--d) depict sensitivities of sea-ice export
658 through Fram Strait in December 1995 to changes in sea-ice thickness
659 12, 24, 36, 48 months back in time. Corresponding sensitivities to
660 ocean surface temperature are depicted in
661 \reffig{4yradjthetalev1}(a--d). The main characteristics is
662 consistency with expected advection of sea-ice over the relevant time
663 scales considered. The general positive pattern means that an
664 increase in sea-ice thickness at location $(x,y)$ and time $t$ will
665 increase sea-ice export through Fram Strait at time $T_e$. Largest
666 distances from Fram Strait indicate fastest sea-ice advection over the
667 time span considered. The ice thickness sensitivities are in close
668 correspondence to ocean surface sentivitites, but of opposite sign.
669 An increase in temperature will incur ice melting, decrease in ice
670 thickness, and therefore decrease in sea-ice export at time $T_e$.
671
672 The picture is fundamentally different and much more complex
673 for sensitivities to ocean temperatures away from the surface.
674 \reffig{4yradjthetalev10??}(a--d) depicts ice export sensitivities to
675 temperatures at roughly 400 m depth.
676 Primary features are the effect of the heat transport of the North
677 Atlantic current which feeds into the West Spitsbergen current,
678 the circulation around Svalbard, and ...
679
680 \begin{figure}[t!]
681 \centerline{
682 \subfigure[{\footnotesize -12 months}]
683 {\includegraphics*[width=0.44\linewidth]{\fpath/run_4yr_ADJheff_arc_lev1_tim072_cmax2.0E+02.eps}}
684 %\includegraphics*[width=.3\textwidth]{H_c.bin_res_100_lev1.pdf}
685 %
686 \subfigure[{\footnotesize -24 months}]
687 {\includegraphics*[width=0.44\linewidth]{\fpath/run_4yr_ADJheff_arc_lev1_tim145_cmax2.0E+02.eps}}
688 }
689
690 \centerline{
691 \subfigure[{\footnotesize
692 -36 months}]
693 {\includegraphics*[width=0.44\linewidth]{\fpath/run_4yr_ADJheff_arc_lev1_tim218_cmax2.0E+02.eps}}
694 %
695 \subfigure[{\footnotesize
696 -48 months}]
697 {\includegraphics*[width=0.44\linewidth]{\fpath/run_4yr_ADJheff_arc_lev1_tim292_cmax2.0E+02.eps}}
698 }
699 \caption{Sensitivity of sea-ice export through Fram Strait in December 2005 to
700 sea-ice thickness at various prior times.
701 \label{fig:4yradjheff}}
702 \end{figure}
703
704
705 \begin{figure}[t!]
706 \centerline{
707 \subfigure[{\footnotesize -12 months}]
708 {\includegraphics*[width=0.44\linewidth]{\fpath/run_4yr_ADJtheta_arc_lev1_tim072_cmax5.0E+01.eps}}
709 %\includegraphics*[width=.3\textwidth]{H_c.bin_res_100_lev1.pdf}
710 %
711 \subfigure[{\footnotesize -24 months}]
712 {\includegraphics*[width=0.44\linewidth]{\fpath/run_4yr_ADJtheta_arc_lev1_tim145_cmax5.0E+01.eps}}
713 }
714
715 \centerline{
716 \subfigure[{\footnotesize
717 -36 months}]
718 {\includegraphics*[width=0.44\linewidth]{\fpath/run_4yr_ADJtheta_arc_lev1_tim218_cmax5.0E+01.eps}}
719 %
720 \subfigure[{\footnotesize
721 -48 months}]
722 {\includegraphics*[width=0.44\linewidth]{\fpath/run_4yr_ADJtheta_arc_lev1_tim292_cmax5.0E+01.eps}}
723 }
724 \caption{Same as \reffig{4yradjheff} but for sea surface temperature
725 \label{fig:4yradjthetalev1}}
726 \end{figure}
727
728
729
730 \section{Discussion and conclusion}
731 \label{sec:concl}
732
733 The story of the paper could be:
734 B-grid ice model + C-grid ocean model does not work properly for us,
735 therefore C-grid ice model with advantages:
736 \begin{enumerate}
737 \item stress coupling simpler (no interpolation required)
738 \item different boundary conditions
739 \item advection schemes carry over trivially from main code
740 \end{enumerate}
741 LSR/EVP solutions are similar with appropriate bcs, evp parameters as
742 a function of forcing time scale (when does VP solution break
743 down). Same for LSR solver, provided that it works (o:
744 Which scheme is more efficient for the resolution/time stepping
745 parameters that we use here. What about other resolutions?
746
747 \paragraph{Acknowledgements}
748 We thank Jinlun Zhang for providing the original B-grid code and many
749 helpful discussions. ML thanks Elizabeth Hunke for multiple explanations.
750
751 \bibliography{bib/journal_abrvs,bib/seaice,bib/genocean,bib/maths,bib/mitgcmuv,bib/fram}
752
753 \end{document}
754
755 %%% Local Variables:
756 %%% mode: latex
757 %%% TeX-master: t
758 %%% End:
759
760
761 A Dynamic-Thermodynamic Sea ice Model for Ocean Climate
762 Estimation on an Arakawa C-Grid
763
764 Introduction
765
766 Ice Model:
767 Dynamics formulation.
768 B-C, LSR, EVP, no-slip, slip
769 parallellization
770 Thermodynamics formulation.
771 0-layer Hibler salinity + snow
772 3-layer Winton
773
774 Idealized tests
775 Funnel Experiments
776 Downstream Island tests
777 B-grid LSR no-slip
778 C-grid LSR no-slip
779 C-grid LSR slip
780 C-grid EVP no-slip
781 C-grid EVP slip
782
783 Arctic Setup
784 Configuration
785 OBCS from cube
786 forcing
787 1/2 and full resolution
788 with a few JFM figs from C-grid LSR no slip
789 ice transport through Canadian Archipelago
790 thickness distribution
791 ice velocity and transport
792
793 Arctic forward sensitivity experiments
794 B-grid LSR no-slip
795 C-grid LSR no-slip
796 C-grid LSR slip
797 C-grid EVP no-slip
798 C-grid EVP slip
799 C-grid LSR no-slip + Winton
800 speed-performance-accuracy (small)
801 ice transport through Canadian Archipelago differences
802 thickness distribution differences
803 ice velocity and transport differences
804
805 Adjoint sensitivity experiment on 1/2-res setup
806 Sensitivity of sea ice volume flow through Fram Strait
807 *** Sensitivity of sea ice volume flow through Canadian Archipelago
808
809 Summary and conluding remarks

  ViewVC Help
Powered by ViewVC 1.1.22