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

Diff of /MITgcm_contrib/articles/ceaice/ceaice.tex

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

revision 1.16 by mlosch, Tue Feb 26 19:14:36 2008 UTC revision 1.23 by dimitri, Thu Aug 14 16:12:41 2008 UTC
# Line 3  Line 3 
3  \documentclass[12pt]{article}  \documentclass[12pt]{article}
4    
5  \usepackage[]{graphicx}  \usepackage[]{graphicx}
6    %\usepackage[draft]{graphicx}
7  \usepackage{subfigure}  \usepackage{subfigure}
8    
9  \usepackage[round,comma]{natbib}  \usepackage[round,comma]{natbib}
# Line 12  Line 13 
13  \newcommand\bmmax{10} \newcommand\hmmax{10}  \newcommand\bmmax{10} \newcommand\hmmax{10}
14  \usepackage{bm}  \usepackage{bm}
15    
16    \usepackage{url}
17    
18  % math abbreviations  % math abbreviations
19  \newcommand{\vek}[1]{\ensuremath{\mathbf{#1}}}  \newcommand{\vek}[1]{\ensuremath{\mathbf{#1}}}
20  \newcommand{\mat}[1]{\ensuremath{\mathbf{#1}}}  \newcommand{\mat}[1]{\ensuremath{\mathbf{#1}}}
# Line 42  Line 45 
45  % commenting scheme  % commenting scheme
46  \newcommand{\ml}[1]{\textsf{\slshape #1}}  \newcommand{\ml}[1]{\textsf{\slshape #1}}
47    
48  \title{A Dynamic-Thermodynamic Sea ice Model for Ocean Climate  \title{A Dynamic-Thermodynamic Sea Ice Model on an Arakawa C-Grid
49    Estimation on an Arakawa C-Grid}    for Ocean Climate Estimation and Sensitivity Studies}
50    
51    %Alternative title suggested by Chris Hill:
52    %\title{A Sea Ice Model Designed for Ocean State Estimation and its
53    %  Application to Studying Sea Ice Model Dynamics in the Canadian Arctic
54    %  Archipelago}
55    
56  \author{Martin Losch, Dimitris Menemenlis, Patrick Heimbach, \\  \author{Martin Losch, Dimitris Menemenlis, Patrick Heimbach, \\
57          Jean-Michel Campin, and Chris Hill}          Jean-Michel Campin, and Chris Hill}
# Line 51  Line 59 
59    
60  \maketitle  \maketitle
61    
62  \begin{abstract}  \input{ceaice_abstract.tex}
63  As part of ongoing efforts to obtain a best possible synthesis of most  
64  available, global-scale, ocean and sea ice data, a dynamic and thermodynamic  \input{ceaice_intro.tex}
65  sea-ice model has been coupled to the Massachusetts Institute of Technology  
66  general circulation model (MITgcm).  Ice mechanics follow a viscous plastic  \input{ceaice_model.tex}
67  rheology and the ice momentum equations are solved numerically using either  
68  line successive relaxation (LSR) or elastic-viscous-plastic (EVP) dynamic  \input{ceaice_forward.tex}
69  models.  Ice thermodynamics are represented using either a zero-heat-capacity  
70  formulation or a two-layer formulation that conserves enthalpy.  The model  \input{ceaice_adjoint.tex}
71  includes prognostic variables for snow and for sea-ice salinity.  The above  
72  sea ice model components were borrowed from current-generation climate models  \input{ceaice_concl.tex}
73  but they were reformulated on an Arakawa C-grid in order to match the MITgcm  
74  oceanic grid and they were modified in many ways to permit efficient and  \appendix
75  accurate automatic differentiation.  This paper describes the MITgcm sea ice  \input{ceaice_appendix.tex}
 model; it presents example Arctic and Antarctic results from a realistic,  
 eddy-permitting, global ocean and sea-ice configuration; it compares B-grid  
 and C-grid dynamic solvers in a regional Arctic configuration; and it presents  
 example results from coupled ocean and sea-ice adjoint-model integrations.  
   
 \end{abstract}  
   
 \section{Introduction}  
 \label{sec:intro}  
   
 The availability of an adjoint model as a powerful research tool  
 complementary to an ocean model was a major design requirement early  
 on in the development of the MIT general circulation model (MITgcm)  
 [Marshall et al. 1997a, Marotzke et al. 1999, Adcroft et al. 2002]. It  
 was recognized that the adjoint model permitted computing the  
 gradients of various scalar-valued model diagnostics, norms or,  
 generally, objective functions with respect to external or independent  
 parameters very efficiently. The information associtated with these  
 gradients is useful in at least two major contexts. First, for state  
 estimation problems, the objective function is the sum of squared  
 differences between observations and model results weighted by the  
 inverse error covariances. The gradient of such an objective function  
 can be used to reduce this measure of model-data misfit to find the  
 optimal model solution in a least-squares sense.  Second, the  
 objective function can be a key oceanographic quantity such as  
 meridional heat or volume transport, ocean heat content or mean  
 surface temperature index. In this case the gradient provides a  
 complete set of sensitivities of this quantity to all independent  
 variables simultaneously. These sensitivities can be used to address  
 the cause of, say, changing net transports accurately.  
   
 References to existing sea-ice adjoint models, explaining that they are either  
 for simplified configurations, for ice-only studies, or for short-duration  
 studies to motivate the present work.  
   
 Traditionally, probably for historical reasons and the ease of  
 treating the Coriolis term, most standard sea-ice models are  
 discretized on Arakawa-B-grids \citep[e.g.,][]{hibler79, harder99,  
   kreyscher00, zhang98, hunke97}. From the perspective of coupling a  
 sea ice-model to a C-grid ocean model, the exchange of fluxes of heat  
 and fresh-water pose no difficulty for a B-grid sea-ice model  
 \citep[e.g.,][]{timmermann02a}. However, surface stress is defined at  
 velocities points and thus needs to be interpolated between a B-grid  
 sea-ice model and a C-grid ocean model. Smoothing implicitly  
 associated with this interpolation may mask grid scale noise and may  
 contribute to stabilizing the solution. On the other hand, by  
 smoothing the stress signals are damped which could lead to reduced  
 variability of the system. By choosing a C-grid for the sea-ice model,  
 we circumvent this difficulty altogether and render the stress  
 coupling as consistent as the buoyancy coupling.  
   
 A further advantage of the C-grid formulation is apparent in narrow  
 straits. In the limit of only one grid cell between coasts there is no  
 flux allowed for a B-grid (with no-slip lateral boundary counditions),  
 and models have used topographies artificially widened straits to  
 avoid this problem \citep{holloway07}. The C-grid formulation on the  
 other hand allows a flux of sea-ice through narrow passages if  
 free-slip along the boundaries is allowed. We demonstrate this effect  
 in the Candian archipelago.  
   
 Talk about problems that make the sea-ice-ocean code very sensitive and  
 changes in the code that reduce these sensitivities.  
   
 This paper describes the MITgcm sea ice  
 model; it presents example Arctic and Antarctic results from a realistic,  
 eddy-permitting, global ocean and sea-ice configuration; it compares B-grid  
 and C-grid dynamic solvers in a regional Arctic configuration; and it presents  
 example results from coupled ocean and sea-ice adjoint-model integrations.  
   
 \section{Model}  
 \label{sec:model}  
   
 \subsection{Dynamics}  
 \label{sec:dynamics}  
   
 The momentum equation of the sea-ice model is  
 \begin{equation}  
   \label{eq:momseaice}  
   m \frac{D\vek{u}}{Dt} = -mf\vek{k}\times\vek{u} + \vtau_{air} +  
   \vtau_{ocean} - m \nabla{\phi(0)} + \vek{F},  
 \end{equation}  
 where $m=m_{i}+m_{s}$ is the ice and snow mass per unit area;  
 $\vek{u}=u\vek{i}+v\vek{j}$ is the ice velocity vector;  
 $\vek{i}$, $\vek{j}$, and $\vek{k}$ are unit vectors in the $x$, $y$, and $z$  
 directions, respectively;  
 $f$ is the Coriolis parameter;  
 $\vtau_{air}$ and $\vtau_{ocean}$ are the wind-ice and ocean-ice stresses,  
 respectively;  
 $g$ is the gravity accelation;  
 $\nabla\phi(0)$ is the gradient (or tilt) of the sea surface height;  
 $\phi(0) = g\eta + p_{a}/\rho_{0}$ is the sea surface height potential  
 in response to ocean dynamics ($g\eta$) and to atmospheric pressure  
 loading ($p_{a}/\rho_{0}$, where $\rho_{0}$ is a reference density);  
 and $\vek{F}=\nabla\cdot\sigma$ is the divergence of the internal ice stress  
 tensor $\sigma_{ij}$.  
 When using the rescaled vertical coordinate system, z$^\ast$, of  
 \citet{cam08}, $\phi(0)$ also includes a term due to snow and ice  
 loading, $mg/\rho_{0}$.  
 Advection of sea-ice momentum is neglected. The wind and ice-ocean stress  
 terms are given by  
 \begin{align*}  
   \vtau_{air}   = & \rho_{air}  C_{air}   |\vek{U}_{air}  -\vek{u}|  
                    R_{air}  (\vek{U}_{air}  -\vek{u}), \\  
   \vtau_{ocean} = & \rho_{ocean}C_{ocean} |\vek{U}_{ocean}-\vek{u}|  
                    R_{ocean}(\vek{U}_{ocean}-\vek{u}), \\  
 \end{align*}  
 where $\vek{U}_{air/ocean}$ are the surface winds of the atmosphere  
 and surface currents of the ocean, respectively; $C_{air/ocean}$ are  
 air and ocean drag coefficients; $\rho_{air/ocean}$ are reference  
 densities; and $R_{air/ocean}$ are rotation matrices that act on the  
 wind/current vectors.  
   
 For an isotropic system the stress tensor $\sigma_{ij}$ ($i,j=1,2$) can  
 be related to the ice strain rate and strength by a nonlinear  
 viscous-plastic (VP) constitutive law \citep{hibler79, zhang98}:  
 \begin{equation}  
   \label{eq:vpequation}  
   \sigma_{ij}=2\eta(\dot{\epsilon}_{ij},P)\dot{\epsilon}_{ij}  
   + \left[\zeta(\dot{\epsilon}_{ij},P) -  
     \eta(\dot{\epsilon}_{ij},P)\right]\dot{\epsilon}_{kk}\delta_{ij}    
   - \frac{P}{2}\delta_{ij}.  
 \end{equation}  
 The ice strain rate is given by  
 \begin{equation*}  
   \dot{\epsilon}_{ij} = \frac{1}{2}\left(  
     \frac{\partial{u_{i}}}{\partial{x_{j}}} +  
     \frac{\partial{u_{j}}}{\partial{x_{i}}}\right).  
 \end{equation*}  
 The maximum ice pressure $P_{\max}$, a measure of ice strength, depends on  
 both thickness $h$ and compactness (concentration) $c$:  
 \begin{equation}  
   P_{\max} = P^{*}c\,h\,e^{[C^{*}\cdot(1-c)]},  
 \label{eq:icestrength}  
 \end{equation}  
 with the constants $P^{*}$ and $C^{*}$. The nonlinear bulk and shear  
 viscosities $\eta$ and $\zeta$ are functions of ice strain rate  
 invariants and ice strength such that the principal components of the  
 stress lie on an elliptical yield curve with the ratio of major to  
 minor axis $e$ equal to $2$; they are given by:  
 \begin{align*}  
   \zeta =& \min\left(\frac{P_{\max}}{2\max(\Delta,\Delta_{\min})},  
    \zeta_{\max}\right) \\  
   \eta =& \frac{\zeta}{e^2} \\  
   \intertext{with the abbreviation}  
   \Delta = & \left[  
     \left(\dot{\epsilon}_{11}^2+\dot{\epsilon}_{22}^2\right)  
     (1+e^{-2}) +  4e^{-2}\dot{\epsilon}_{12}^2 +  
     2\dot{\epsilon}_{11}\dot{\epsilon}_{22} (1-e^{-2})  
   \right]^{-\frac{1}{2}}  
 \end{align*}  
 The bulk viscosities are bounded above by imposing both a minimum  
 $\Delta_{\min}=10^{-11}\text{\,s}^{-1}$ (for numerical reasons) and a  
 maximum $\zeta_{\max} = P_{\max}/\Delta^*$, where  
 $\Delta^*=(5\times10^{12}/2\times10^4)\text{\,s}^{-1}$. For stress  
 tensor computation the replacement pressure $P = 2\,\Delta\zeta$  
 \citep{hibler95} is used so that the stress state always lies on the  
 elliptic yield curve by definition.  
   
 In the so-called truncated ellipse method the shear viscosity $\eta$  
 is capped to suppress any tensile stress \citep{hibler97, geiger98}:  
 \begin{equation}  
   \label{eq:etatem}  
   \eta = \min\left(\frac{\zeta}{e^2},  
   \frac{\frac{P}{2}-\zeta(\dot{\epsilon}_{11}+\dot{\epsilon}_{22})}  
   {\sqrt{(\dot{\epsilon}_{11}+\dot{\epsilon}_{22})^2  
       +4\dot{\epsilon}_{12}^2}}\right).  
 \end{equation}  
   
 In the current implementation, the VP-model is integrated with the  
 semi-implicit line successive over relaxation (LSOR)-solver of  
 \citet{zhang98}, which allows for long time steps that, in our case,  
 are limited by the explicit treatment of the Coriolis term. The  
 explicit treatment of the Coriolis term does not represent a severe  
 limitation because it restricts the time step to approximately the  
 same length as in the ocean model where the Coriolis term is also  
 treated explicitly.  
   
 \citet{hunke97}'s introduced an elastic contribution to the strain  
 rate in order to regularize Eq.\refeq{vpequation} in such a way that  
 the resulting elastic-viscous-plastic (EVP) and VP models are  
 identical at steady state,  
 \begin{equation}  
   \label{eq:evpequation}  
   \frac{1}{E}\frac{\partial\sigma_{ij}}{\partial{t}} +  
   \frac{1}{2\eta}\sigma_{ij}  
   + \frac{\eta - \zeta}{4\zeta\eta}\sigma_{kk}\delta_{ij}    
   + \frac{P}{4\zeta}\delta_{ij}  
   = \dot{\epsilon}_{ij}.  
 \end{equation}  
 %In the EVP model, equations for the components of the stress tensor  
 %$\sigma_{ij}$ are solved explicitly. Both model formulations will be  
 %used and compared the present sea-ice model study.  
 The EVP-model uses an explicit time stepping scheme with a short  
 timestep. According to the recommendation of \citet{hunke97}, the  
 EVP-model is stepped forward in time 120 times within the physical  
 ocean model time step (although this parameter is under debate), to  
 allow for elastic waves to disappear.  Because the scheme does not  
 require a matrix inversion it is fast in spite of the small timestep  
 \citep{hunke97}. For completeness, we repeat the equations for the  
 components of the stress tensor $\sigma_{1} =  
 \sigma_{11}+\sigma_{22}$, $\sigma_{2}= \sigma_{11}-\sigma_{22}$, and  
 $\sigma_{12}$. Introducing the divergence $D_D =  
 \dot{\epsilon}_{11}+\dot{\epsilon}_{22}$, and the horizontal tension  
 and shearing strain rates, $D_T =  
 \dot{\epsilon}_{11}-\dot{\epsilon}_{22}$ and $D_S =  
 2\dot{\epsilon}_{12}$, respectively, and using the above abbreviations,  
 the equations can be written as:  
 \begin{align}  
   \label{eq:evpstresstensor1}  
   \frac{\partial\sigma_{1}}{\partial{t}} + \frac{\sigma_{1}}{2T} +  
   \frac{P}{2T} &= \frac{P}{2T\Delta} D_D \\  
   \label{eq:evpstresstensor2}  
   \frac{\partial\sigma_{2}}{\partial{t}} + \frac{\sigma_{2} e^{2}}{2T}  
   &= \frac{P}{2T\Delta} D_T \\  
   \label{eq:evpstresstensor12}  
   \frac{\partial\sigma_{12}}{\partial{t}} + \frac{\sigma_{12} e^{2}}{2T}  
   &= \frac{P}{4T\Delta} D_S  
 \end{align}  
 Here, the elastic parameter $E$ is redefined in terms of a damping timescale  
 $T$ for elastic waves \[E=\frac{\zeta}{T}.\]  
 $T=E_{0}\Delta{t}$ with the tunable parameter $E_0<1$ and  
 the external (long) timestep $\Delta{t}$. \citet{hunke97} recommend  
 $E_{0} = \frac{1}{3}$.  
   
 For details of the spatial discretization, the reader is referred to  
 \citet{zhang98, zhang03}. Our discretization differs only (but  
 importantly) in the underlying grid, namely the Arakawa C-grid, but is  
 otherwise straightforward. The EVP model, in particular, is discretized  
 naturally on the C-grid with $\sigma_{1}$ and $\sigma_{2}$ on the  
 center points and $\sigma_{12}$ on the corner (or vorticity) points of  
 the grid. With this choice all derivatives are discretized as central  
 differences and averaging is only involved in computing $\Delta$ and  
 $P$ at vorticity points.  
   
 For a general curvilinear grid, one needs in principle to take metric  
 terms into account that arise in the transformation of a curvilinear  
 grid on the sphere. For now, however, we can neglect these metric  
 terms because they are very small on the \ml{[modify following  
   section3:] cubed sphere grids used in this paper; in particular,  
 only near the edges of the cubed sphere grid, we expect them to be  
 non-zero, but these edges are at approximately 35\degS\ or 35\degN\  
 which are never covered by sea-ice in our simulations.  Everywhere  
 else the coordinate system is locally nearly cartesian.}  However, for  
 last-glacial-maximum or snowball-earth-like simulations the question  
 of metric terms needs to be reconsidered.  Either, one includes these  
 terms as in \citet{zhang03}, or one finds a vector-invariant  
 formulation for the sea-ice internal stress term that does not require  
 any metric terms, as it is done in the ocean dynamics of the MITgcm  
 \citep{adcroft04:_cubed_sphere}.  
   
 Lateral boundary conditions are naturally ``no-slip'' for B-grids, as  
 the tangential velocities points lie on the boundary. For C-grids, the  
 lateral boundary condition for tangential velocities is realized via  
 ``ghost points'', allowing alternatively no-slip or free-slip  
 conditions. In ocean models free-slip boundary conditions in  
 conjunction with piecewise-constant (``castellated'') coastlines have  
 been shown to reduce in effect to no-slip boundary conditions  
 \citep{adcroft98:_slippery_coast}; for sea-ice models the effects of  
 lateral boundary conditions have not yet been studied.  
   
 Moving sea ice exerts a stress on the ocean which is the opposite of  
 the stress $\vtau_{ocean}$ in Eq.\refeq{momseaice}. This stess is  
 applied directly to the surface layer of the ocean model. An  
 alternative ocean stress formulation is given by \citet{hibler87}.  
 Rather than applying $\vtau_{ocean}$ directly, the stress is derived  
 from integrating over the ice thickness to the bottom of the oceanic  
 surface layer. In the resulting equation for the \emph{combined}  
 ocean-ice momentum, the interfacial stress cancels and the total  
 stress appears as the sum of windstress and divergence of internal ice  
 stresses: $\delta(z) (\vtau_{air} + \vek{F})/\rho_0$, \citep[see also  
 Eq.\,2 of][]{hibler87}. The disadvantage of this formulation is that  
 now the velocity in the surface layer of the ocean that is used to  
 advect tracers, is really an average over the ocean surface  
 velocity and the ice velocity leading to an inconsistency as the ice  
 temperature and salinity are different from the oceanic variables.  
   
 Sea ice distributions are characterized by sharp gradients and edges.  
 For this reason, we employ positive, multidimensional 2nd and 3rd-order  
 advection scheme with flux limiter \citep{roe85, hundsdorfer94} for the  
 thermodynamic variables discussed in the next section.  
   
 \subparagraph{boundary conditions: no-slip, free-slip, half-slip}  
   
 \begin{itemize}  
 \item transition from existing B-Grid to C-Grid  
 \item boundary conditions: no-slip, free-slip, half-slip  
 \item fancy (multi dimensional) advection schemes  
 \item VP vs.\ EVP \citep{hunke97}  
 \item ocean stress formulation \citep{hibler87}  
 \end{itemize}  
   
 \subsection{Thermodynamics}  
 \label{sec:thermodynamics}  
   
 In the original formulation the sea ice model \citep{menemenlis05}  
 uses simple thermodynamics following the appendix of  
 \citet{semtner76}. This formulation does not allow storage of heat  
 (heat capacity of ice is zero, and this type of model is often refered  
 to as a ``zero-layer'' model). Upward heat flux is parameterized  
 assuming a linear temperature profile and together with a constant ice  
 conductivity. It is expressed as $(K/h)(T_{w}-T_{0})$, where $K$ is  
 the ice conductivity, $h$ the ice thickness, and $T_{w}-T_{0}$ the  
 difference between water and ice surface temperatures. The surface  
 heat budget is computed in a similar way to that of  
 \citet{parkinson79} and \citet{manabe79}.  
   
 There is considerable doubt about the reliability of such a simple  
 thermodynamic model---\citet{semtner84} found significant errors in  
 phase (one month lead) and amplitude ($\approx$50\%\,overestimate) in  
 such models---, so that today many sea ice models employ more complex  
 thermodynamics. A popular thermodynamics model of \citet{winton00} is  
 based on the 3-layer model of \citet{semtner76} and treats brine  
 content by means of enthalphy conservation. This model requires in  
 addition to ice-thickness and compactness (fractional area) additional  
 state variables to be advected by ice velocities, namely enthalphy of  
 the two ice layers and the thickness of the overlying snow layer.  
 \ml{[Jean-Michel, your turf: ]Care must be taken in advecting these  
   quantities in order to ensure conservation of enthalphy. Currently  
   this can only be accomplished with a 2nd-order advection scheme with  
   flux limiter \citep{roe85}.}  
   
   
 \subsection{C-grid}  
 \begin{itemize}  
 \item no-slip vs. free-slip for both lsr and evp;  
   "diagnostics" to look at and use for comparison  
   \begin{itemize}  
   \item ice transport through Fram Strait/Denmark Strait/Davis  
     Strait/Bering strait (these are general)  
   \item ice transport through narrow passages, e.g.\ Nares-Strait  
   \end{itemize}  
 \item compare different advection schemes (if lsr turns out to be more  
   effective, then with lsr otherwise I prefer evp), eg.  
   \begin{itemize}  
   \item default 2nd-order with diff1=0.002  
   \item 1st-order upwind with diff1=0.  
   \item DST3FL (SEAICEadvScheme=33 with diff1=0., should work, works for me)  
   \item 2nd-order wit flux limiter (SEAICEadvScheme=77 with diff1=0.)  
   \end{itemize}  
   That should be enough. Here, total ice mass and location of ice edge  
   is interesting. However, this comparison can be done in an idealized  
   domain, may not require full Arctic Domain?  
 \item  
 Do a little study on the parameters of LSR and EVP  
 \begin{enumerate}  
 \item convergence of LSR, how many iterations do you need to get a  
   true elliptic yield curve  
 \item vary deltaTevp and the relaxation parameter for EVP and see when  
   the EVP solution breaks down (relative to the forcing time scale).  
   For this, it is essential that the evp solver gives use "stripeless"  
   solutions, that is your dtevp = 1sec solutions/or 10sec solutions  
   with SEAICE\_evpDampC = 615.  
 \end{enumerate}  
 \end{itemize}  
   
 \section{Forward sensitivity experiments}  
 \label{sec:forward}  
   
 A second series of forward sensitivity experiments have been carried out on an  
 Arctic Ocean domain with open boundaries.  Once again the objective is to  
 compare the old B-grid LSR dynamic solver with the new C-grid LSR and EVP  
 solvers.  One additional experiment is carried out to illustrate the  
 differences between the two main options for sea ice thermodynamics in the MITgcm.  
   
 \subsection{Arctic Domain with Open Boundaries}  
 \label{sec:arctic}  
   
 The Arctic domain of integration is illustrated in Fig.~\ref{fig:arctic1}.  It  
 is carved out from, and obtains open boundary conditions from, the  
 global cubed-sphere configuration of the Estimating the Circulation  
 and Climate of the Ocean, Phase II (ECCO2) project  
 \citet{menemenlis05}.  The domain size is 420 by 384 grid boxes  
 horizontally with mean horizontal grid spacing of 18 km.  
   
 \begin{figure}  
 %\centerline{{\includegraphics*[width=0.44\linewidth]{\fpath/arctic1.eps}}}  
 \caption{Bathymetry of Arctic Domain.\label{fig:arctic1}}  
 \end{figure}  
   
 There are 50 vertical levels ranging in thickness from 10 m near the surface  
 to approximately 450 m at a maximum model depth of 6150 m. Bathymetry is from  
 the National Geophysical Data Center (NGDC) 2-minute gridded global relief  
 data (ETOPO2) and the model employs the partial-cell formulation of  
 \citet{adcroft97:_shaved_cells}, which permits accurate representation of the bathymetry. The  
 model is integrated in a volume-conserving configuration using a finite volume  
 discretization with C-grid staggering of the prognostic variables. In the  
 ocean, the non-linear equation of state of \citet{jackett95}.  The ocean model is  
 coupled to a sea-ice model described hereinabove.    
   
 This particular ECCO2 simulation is initialized from rest using the  
 January temperature and salinity distribution from the World Ocean  
 Atlas 2001 (WOA01) [Conkright et al., 2002] and it is integrated for  
 32 years prior to the 1996--2001 period discussed in the study. Surface  
 boundary conditions are from the National Centers for Environmental  
 Prediction and the National Center for Atmospheric Research  
 (NCEP/NCAR) atmospheric reanalysis [Kistler et al., 2001]. Six-hourly  
 surface winds, temperature, humidity, downward short- and long-wave  
 radiations, and precipitation are converted to heat, freshwater, and  
 wind stress fluxes using the \citet{large81, large82} bulk formulae.  
 Shortwave radiation decays exponentially as per Paulson and Simpson  
 [1977]. Additionally the time-mean river run-off from Large and Nurser  
 [2001] is applied and there is a relaxation to the monthly-mean  
 climatological sea surface salinity values from WOA01 with a  
 relaxation time scale of 3 months. Vertical mixing follows  
 \citet{large94} with background vertical diffusivity of  
 $1.5\times10^{-5}\text{\,m$^{2}$\,s$^{-1}$}$ and viscosity of  
 $10^{-3}\text{\,m$^{2}$\,s$^{-1}$}$. A third order, direct-space-time  
 advection scheme with flux limiter is employed \citep{hundsdorfer94}  
 and there is no explicit horizontal diffusivity. Horizontal viscosity  
 follows \citet{lei96} but  
 modified to sense the divergent flow as per Fox-Kemper and Menemenlis  
 [in press].  Shortwave radiation decays exponentially as per Paulson  
 and Simpson [1977].  Additionally, the time-mean runoff of Large and  
 Nurser [2001] is applied near the coastline and, where there is open  
 water, there is a relaxation to monthly-mean WOA01 sea surface  
 salinity with a time constant of 45 days.  
   
 Open water, dry  
 ice, wet ice, dry snow, and wet snow albedo are, respectively, 0.15, 0.85,  
 0.76, 0.94, and 0.8.  
   
 \begin{itemize}  
 \item Configuration  
 \item OBCS from cube  
 \item forcing  
 \item 1/2 and full resolution  
 \item with a few JFM figs from C-grid LSR no slip  
   ice transport through Canadian Archipelago  
   thickness distribution  
   ice velocity and transport  
 \end{itemize}  
   
 \begin{itemize}  
 \item Arctic configuration  
 \item ice transport through straits and near boundaries  
 \item focus on narrow straits in the Canadian Archipelago  
 \end{itemize}  
   
 \begin{itemize}  
 \item B-grid LSR no-slip  
 \item C-grid LSR no-slip  
 \item C-grid LSR slip  
 \item C-grid EVP no-slip  
 \item C-grid EVP slip  
 \item C-grid LSR + TEM (truncated ellipse method, no tensile stress, new flag)  
 \item C-grid LSR no-slip + Winton  
 \item  speed-performance-accuracy (small)  
   ice transport through Canadian Archipelago differences  
   thickness distribution differences  
   ice velocity and transport differences  
 \end{itemize}  
   
 We anticipate small differences between the different models due to:  
 \begin{itemize}  
 \item advection schemes: along the ice-edge and regions with large  
   gradients  
 \item C-grid: less transport through narrow straits for no slip  
   conditons, more for free slip  
 \item VP vs.\ EVP: speed performance, accuracy?  
 \item ocean stress: different water mass properties beneath the ice  
 \end{itemize}  
   
 \section{Adjoint sensiivities of the MITsim}  
   
 \subsection{The adjoint of MITsim}  
   
 The ability to generate tangent linear and adjoint model components  
 of the MITsim has been a main design task.  
 For the ocean the adjoint capability has proven to be an  
 invaluable tool for sensitivity analysis as well as state estimation.  
 In short, the adjoint enables very efficient computation of the gradient  
 of scalar-valued model diagnostics (called cost function or objective function)  
 with respect to many model "variables".  
 These variables can be two- or three-dimensional fields of initial  
 conditions, model parameters such as mixing coefficients, or  
 time-varying surface or lateral (open) boundary conditions.  
 When combined, these variables span a potentially high-dimensional  
 (e.g. O(10$^8$)) so-called control space. Performing parameter perturbations  
 to assess model sensitivities quickly becomes prohibitive at these scales.  
 Alternatively, (time-varying) sensitivities of the objective function  
 to any element of the  control space can be computed very efficiently in  
 one single adjoint  
 model integration, provided an efficient adjoint model is available.  
   
 [REFERENCES]  
   
   
 The adjoint operator (ADM) is the transpose of the tangent linear operator (TLM)  
 of the full (in general nonlinear) forward model, i.e. the MITsim.  
 The TLM maps perturbations of elements of the control space  
 (e.g. initial ice thickness distribution)  
 via the model Jacobian  
 to a perturbation in the objective function  
 (e.g. sea-ice export at the end of the integration interval).  
 \textit{Tangent} linearity ensures that the derivatives are evaluated  
 with respect to the underlying model trajectory at each point in time.  
 This is crucial for nonlinear trajectories and the presence of different  
 regimes (e.g. effect of the seaice growth term at or away from the  
 freezing point of the ocean surface).  
 Ensuring tangent linearity can be easily achieved by integrating  
 the full model in sync with the TLM to provide the underlying model state.  
 Ensuring \textit{tangent} adjoints is equally crucial, but much more  
 difficult to achieve because of the reverse nature of the integration:  
 the adjoint accumulates sensitivities backward in time,  
 starting from a unit perturbation of the objective function.  
 The adjoint model requires the model state in reverse order.  
 This presents one of the major complications in deriving an  
 exact, i.e. \textit{tangent} adjoint model.  
   
 Following closely the development and maintenance of TLM and ADM  
 components of the MITgcm we have relied heavily on the  
 autmomatic differentiation (AD) tool  
 "Transformation of Algorithms in Fortran" (TAF)  
 developed by Fastopt (Giering and Kaminski, 1998)  
 to derive TLM and ADM code of the MITsim.  
 Briefly, the nonlinear parent model is fed to the AD tool which produces  
 derivative code for the specified control space and objective function.  
 Following this approach has (apart from its evident success)  
 several advantages:  
 (1) the adjoint model is the exact adjoint operator of the parent model,  
 (2) the adjoint model can be kept up to date with respect to ongoing  
 development of the parent model, and adjustments to the parent model  
 to extend the automatically generated adjoint are incremental changes  
 only, rather than extensive re-developments,  
 (3) the parallel structure of the parent model is preserved  
 by the adjoint model, ensuring efficient use in high performance  
 computing environments.  
   
 Some initial code adjustments are required to support dependency analysis  
 of the flow reversal and certain language limitations which may lead  
 to irreducible flow graphs (e.g. GOTO statements).  
 The problem of providing the required model state in reverse order  
 at the time of evaluating nonlinear or conditional  
 derivatives is solved via balancing  
 storing vs. recomputation of the model state in a multi-level  
 checkpointing loop.  
 Again, an initial code adjustment is required to support TAFs  
 checkpointing capability.  
 The code adjustments are sufficiently simple so as not to cause  
 major limitations to the full nonlinear parent model.  
 Once in place, an adjoint model of a new model configuration  
 may be derived in about 10 minutes.  
   
 [HIGHLIGHT COUPLED NATURE OF THE ADJOINT!]  
   
 \subsection{Special considerations}  
   
 * growth term(?)  
   
 * small active denominators  
   
 * dynamic solver (implicit function theorem)  
   
 * approximate adjoints  
   
   
 \subsection{An example: sensitivities of sea-ice export through Fram Strait}  
   
 We demonstrate the power of the adjoint method  
 in the context of investigating sea-ice export sensitivities through Fram Strait  
 (for details of this study see Heimbach et al., 2007).  
 %\citep[for details of this study see][]{heimbach07}. %Heimbach et al., 2007).  
 The domain chosen is a coarsened version of the Arctic face of the  
 high-resolution cubed-sphere configuration of the ECCO2 project  
 \citep[see][]{menemenlis05}. It covers the entire Arctic,  
 extends into the North Pacific such as to cover the entire  
 ice-covered regions, and comprises parts of the North Atlantic  
 down to XXN to enable analysis of remote influences of the  
 North Atlantic current to sea-ice variability and export.  
 The horizontal resolution varies between XX and YY km  
 with 50 unevenly spaced vertical levels.  
 The adjoint models run efficiently on 80 processors  
 (benchmarks have been performed both on an SGI Altix as well as an  
 IBM SP5 at NASA/ARC).  
   
 Following a 1-year spinup, the model has been integrated for four  
 years between 1992 and 1995. It is forced using realistic 6-hourly  
 NCEP/NCAR atmospheric state variables. Over the open ocean these are  
 converted into air-sea fluxes via the bulk formulae of  
 \citet{large04}.  Derivation of air-sea fluxes in the presence of  
 sea-ice is handled by the ice model as described in \refsec{model}.  
 The objective function chosen is sea-ice export through Fram Strait  
 computed for December 1995.  The adjoint model computes sensitivities  
 to sea-ice export back in time from 1995 to 1992 along this  
 trajectory.  In principle all adjoint model variable (i.e., Lagrange  
 multipliers) of the coupled ocean/sea-ice model are available to  
 analyze the transient sensitivity behaviour of the ocean and sea-ice  
 state.  Over the open ocean, the adjoint of the bulk formula scheme  
 computes sensitivities to the time-varying atmospheric state.  Over  
 ice-covered parts, the sea-ice adjoint converts surface ocean  
 sensitivities to atmospheric sensitivities.  
   
 \reffig{4yradjheff}(a--d) depict sensitivities of sea-ice export  
 through Fram Strait in December 1995 to changes in sea-ice thickness  
 12, 24, 36, 48 months back in time. Corresponding sensitivities to  
 ocean surface temperature are depicted in  
 \reffig{4yradjthetalev1}(a--d).  The main characteristics is  
 consistency with expected advection of sea-ice over the relevant time  
 scales considered.  The general positive pattern means that an  
 increase in sea-ice thickness at location $(x,y)$ and time $t$ will  
 increase sea-ice export through Fram Strait at time $T_e$.  Largest  
 distances from Fram Strait indicate fastest sea-ice advection over the  
 time span considered.  The ice thickness sensitivities are in close  
 correspondence to ocean surface sentivitites, but of opposite sign.  
 An increase in temperature will incur ice melting, decrease in ice  
 thickness, and therefore decrease in sea-ice export at time $T_e$.  
   
 The picture is fundamentally different and much more complex  
 for sensitivities to ocean temperatures away from the surface.  
 \reffig{4yradjthetalev10??}(a--d) depicts ice export sensitivities to  
 temperatures at roughly 400 m depth.  
 Primary features are the effect of the heat transport of the North  
 Atlantic current which feeds into the West Spitsbergen current,  
 the circulation around Svalbard, and ...  
   
 \begin{figure}[t!]  
 \centerline{  
 \subfigure[{\footnotesize -12 months}]  
 {\includegraphics*[width=0.44\linewidth]{\fpath/run_4yr_ADJheff_arc_lev1_tim072_cmax2.0E+02.eps}}  
 %\includegraphics*[width=.3\textwidth]{H_c.bin_res_100_lev1.pdf}  
 %  
 \subfigure[{\footnotesize -24 months}]  
 {\includegraphics*[width=0.44\linewidth]{\fpath/run_4yr_ADJheff_arc_lev1_tim145_cmax2.0E+02.eps}}  
 }  
   
 \centerline{  
 \subfigure[{\footnotesize  
 -36 months}]  
 {\includegraphics*[width=0.44\linewidth]{\fpath/run_4yr_ADJheff_arc_lev1_tim218_cmax2.0E+02.eps}}  
 %  
 \subfigure[{\footnotesize  
 -48 months}]  
 {\includegraphics*[width=0.44\linewidth]{\fpath/run_4yr_ADJheff_arc_lev1_tim292_cmax2.0E+02.eps}}  
 }  
 \caption{Sensitivity of sea-ice export through Fram Strait in December 2005 to  
 sea-ice thickness at various prior times.  
 \label{fig:4yradjheff}}  
 \end{figure}  
   
   
 \begin{figure}[t!]  
 \centerline{  
 \subfigure[{\footnotesize -12 months}]  
 {\includegraphics*[width=0.44\linewidth]{\fpath/run_4yr_ADJtheta_arc_lev1_tim072_cmax5.0E+01.eps}}  
 %\includegraphics*[width=.3\textwidth]{H_c.bin_res_100_lev1.pdf}  
 %  
 \subfigure[{\footnotesize -24 months}]  
 {\includegraphics*[width=0.44\linewidth]{\fpath/run_4yr_ADJtheta_arc_lev1_tim145_cmax5.0E+01.eps}}  
 }  
   
 \centerline{  
 \subfigure[{\footnotesize  
 -36 months}]  
 {\includegraphics*[width=0.44\linewidth]{\fpath/run_4yr_ADJtheta_arc_lev1_tim218_cmax5.0E+01.eps}}  
 %  
 \subfigure[{\footnotesize  
 -48 months}]  
 {\includegraphics*[width=0.44\linewidth]{\fpath/run_4yr_ADJtheta_arc_lev1_tim292_cmax5.0E+01.eps}}  
 }  
 \caption{Same as \reffig{4yradjheff} but for sea surface temperature  
 \label{fig:4yradjthetalev1}}  
 \end{figure}  
   
   
   
 \section{Discussion and conclusion}  
 \label{sec:concl}  
   
 The story of the paper could be:  
 B-grid ice model + C-grid ocean model does not work properly for us,  
 therefore C-grid ice model  with advantages:  
 \begin{enumerate}  
 \item stress coupling simpler (no interpolation required)  
 \item different boundary conditions  
 \item advection schemes carry over trivially from main code  
 \end{enumerate}  
 LSR/EVP solutions are similar with appropriate bcs, evp parameters as  
 a function of forcing time scale (when does VP solution break  
 down). Same for LSR solver, provided that it works (o:  
 Which scheme is more efficient for the resolution/time stepping  
 parameters that we use here. What about other resolutions?  
76    
77  \paragraph{Acknowledgements}  \paragraph{Acknowledgements}
78  We thank Jinlun Zhang for providing the original B-grid code and many  We thank Jinlun Zhang for providing the original B-grid code and many
79  helpful discussions. ML thanks Elizabeth Hunke for multiple explanations.  helpful discussions. ML thanks Elizabeth Hunke for multiple explanations.
80    
81  \bibliography{bib/journal_abrvs,bib/seaice,bib/genocean,bib/maths,bib/mitgcmuv,bib/fram}  This work is a contribution to Estimating the Circulation and Climate of the
82    Ocean, Phase II (ECCO2).  The ECCO2 project (http://ecco2.org/) is sponsored
83    by the NASA Modeling Analysis and Prediction (MAP) program.  D. Menemenlis
84    carried out this work at the Jet Propulsion Laboratory, California Institute
85    of Technology under contract with the National Aeronautics and Space
86    Administration.
87    
88    \bibliography{bib/journal_abrvs,bib/seaice,bib/genocean,bib/maths,bib/mitgcmuv,bib/fram,bib/mit_biblio}
89    
90  \end{document}  \end{document}
91    

Legend:
Removed from v.1.16  
changed lines
  Added in v.1.23

  ViewVC Help
Powered by ViewVC 1.1.22