1. Introduction
A model of sea ice dynamics predicts the movement of the ice pack based on winds, ocean currents, and a model of the material strength of the ice. Nonuniform motion of the ice is responsible for the thickness and extent of the ice pack, which in turn influences the exchange of energy between the atmosphere and polar oceans. The dynamic characteristics of sea ice thereby play an essential role in climate-related processes of the ocean and atmosphere.
Many models have been developed to describe the ice dynamics. Some early studies focused on free drift descriptions with no ice interaction (Felzenbaum 1961; Bryan et al. 1975; Manabe et al. 1979; Parkinson and Washington 1979); others included more complex sea ice rheologies, treating the ice as a Newtonian viscous fluid (Campbell 1965), a linear viscous fluid (Hibler 1974; Hibler and Tucker 1979), or a plastic material. The Arctic Ice Dynamics Joint Experiment (AIDJEX) in the 1970s proposed an elastic–plastic rheology for the sea ice pack (Coon et al. 1974), and several other nonlinear plastic rheologies have been studied since then (e.g., Pritchard et al. 1977; Flato and Hibler 1992; Ip et al. 1991). A nonlinear viscous–plastic (VP) rheology proposed by Hibler (1979) has become the standard sea ice dynamics model and the basis for many recent sea ice studies.
The VP model suffers from numerical difficulties related to the enormous range of effective viscosities present in the model and requires large computational resources that become particularly cumbersome when the model is coupled to an ocean or atmophere model (Hibler and Bryan 1987; Oberhuber 1993a,b). To avoid the stringent time step restriction for stability of an explicit numerical scheme in regions where the ice is relatively rigid, the model equations are typically solved with implicit methods such as successive overrelaxation (Hibler 1979) and line relaxation (Oberhuber 1993a; Holland et al. 1993). However, these methods suffer from poor convergence characteristics as the mesh resolution is increased. Attempts to overcome the inherent problems of the model have included improved numerical methods as well as simplifications of the model itself. As part of this paper, we present a more efficient implicit numerical method for solving the VP model equations that uses preconditioned conjugate gradients.
Simpler versions of the VP model, such as free drift descriptions with no ice interaction and cavitating fluid models in which the ice has no resistance to shear forces (Nikiforov et al. 1967; Flato and Hibler 1989, 1992), are more tractable numerically, but the model behavior is sensitive to these simplifications (Holland et al. 1993). Likewise, simulations with more complicated rheologies than the standard elliptical yield curve (Hibler 1979), such as teardrop (Coon et al. 1974), sine wave lens (Bratchie 1984), Mohr–Coulomb, and square (Ip et al. 1991) shapes, show that the rheology can have a significant effect on long-term simulations of ice drift (Ip et al. 1991). Since an ice model need only simulate a visco–plastic material at timescales on the order of those imposed by wind forcing (days), we also present a modification of the model, the addition of elastic behavior, that realizes significant gains in numerical efficiency, reduces to the original VP model behavior at long timescales, and is more accurate for transients. Our model avoids the complexities of the early elastic–plastic models (Pritchard 1975; Colony and Pritchard 1975) because the elastic-like behavior is not intended to be physically realistic and is introduced for numerical expediency.
The VP model also suffers from inaccuracies in calculating transient behavior. For example, given daily time steps, the VP model behavior is acceptable only for surface stresses that vary on the order of a week or more. Hibler (1979) states that several time steps are needed between changes in the forcing (he uses 8-day averaged winds with a 1-day time step), and more recently, Stössel et al. (1994) have noted that the sea ice components of some ice–ocean coupled models are slow to converge, especially under daily forcing. The VP numerical model does produce correct transient behavior if the time step is taken sufficiently small, on the order of minutes for 1-day forcing timescales. Our implementation of the elastic–viscous–plastic (EVP) model is more accurate in resolving transients, even using relatively large time steps, and therefore will produce more accurate ice behavior.
The VP ice dynamics model is not well suited to parallel architectures. Implicit methods required for larger time steps typically entail a great deal of communication between processors, making parallel computation less attractive. Therefore, explicit models are generally preferable for parallel implementations. Ip et al. (1991) optimized the VP model for multiprocessor computers using an explicit, Euler time-stepping scheme, but stability requirements of the numerical method severely limited the time step. The new EVP model presented in this paper permits a fully explicit implementation with an acceptably long time step. Its efficiency is compared with three methods of solving the viscous–plastic equations: the preconditioned conjugate gradient method and two relaxation schemes (Hibler 1979; Zhang and Hibler 1996).
The present work is part of an effort to develop a computationally efficient sea ice component for a fully coupled atmosphere–ice–ocean global climate model. The sea ice model, which also includes thermodynamic and transport components, is designed to be compatible with the Parallel Ocean Program (POP), an ocean circulation model developed at Los Alamos National Laboratory for use on massively parallel computers (Smith et al. 1992; Dukowicz et al. 1993, 1994).
2. The ice dynamics model
a. Viscous–plastic model equations
Pack ice typically consists of rigid plates, which may drift freely in areas of relatively open water or be closely packed together in regions of high ice concentration. Although individual ice floes range from tens of meters to several kilometers across, the ice pack is considered to be a highly fractured two-dimensional continuum, to make modeling it tractable (Pritchard 1975; Rothrock 1975b; Hibler 1980; Gray and Morland 1994).
There has been a great deal of disagreement about the relative importance of the various terms in (1) (Parkinson and Washington 1979). The primary components are the air and water stresses, Coriolis force, and ice interaction effects (Hibler 1986); the most predominant of these is wind stress (Coon 1980). Rothrock (1975a) demonstrated through scale analyses that the acceleration term is three orders of magnitude smaller than the stress terms. In contrast to Hibler (1979) and following Oberhuber (1993a), we neglect nonlinear advection, which is at least an order of magnitude smaller than the acceleration term. The ice interaction term is essential in balancing the stresses in much of the ice field (Hibler 1979; Parkinson and Washington 1979; Coon 1980; Hibler 1986), and although they are smaller in magnitude, current and tilt effects are significant over long periods of time (Hibler 1986; Warn-Varnas et al. 1991).
The pressure P, a measure of ice strength, depends on both thickness and compactness: P = P*cHe exp[−c*(1 − c)], where P* and c* are constants given in Table 1. This definition of P is equivalent to the standard formulation of Hibler (1979), because cH is approximately the same as his “equivalent ice thickness” h.
b. Motivation for alternative methods
Most sea ice models, starting with the models developed for the AIDJEX project, agree on a visco–plastic rheology at normal levels of strain rate, differing perhaps in the shape of the yield curve. The ideal visco–plastic rheology, however, becomes singular as the strain rate approaches zero. The AIDJEX model (Coon et al. 1974; Pritchard 1975) regularized this behavior by adopting a rheology that converts to that of an elastic material at small strain rates (an “elastic–plastic” rheology). It is important to realize that such an elastic rheology is physically realistic for ice only at a laboratory scale, but at geophysical scales there is no reason to prefer the elastic regularization to any other closure that ensures that the ice pack behaves as a rigid slab in the singular, small strain rate regime. Unfortunately, the AIDJEX model took the limiting elastic behavior quite literally (Pritchard 1975; Colony and Pritchard 1975), needlessly introducing severe theoretical and numerical complexities. Hibler (1979), on the other hand, realized that what was really needed was a simple regularization that gave sufficiently “rigid” behavior in the singular regime. He introduced a regularization (i.e., a “viscous–plastic” rheology), as described previously, in which the nonlinear viscosity of the visco–plastic rheology was bounded at a very high value such that the limiting behavior was really a very slow creep. This regularization, although simple, has its own severe numerical difficulties, which will be discussed shortly. An elastic formulation, on the other hand, has certain advantages from the numerical point of view when viewed merely as a regularization, as explained below. Thus, ironically, we are led to reintroduce a model that has some resemblance to the original AIDJEX model, but in which the regularization is a simplified elastic model whose parameters are chosen for numerical, rather than physical reasons.
c. The elastic formulation
3. Numerical formulations
In this section we outline our numerical techniques for both the preconditioned conjugate gradient method and the explicit elastic–viscous–plastic method. The spatial discretization is specialized for a generalized orthogonal B-grid as in Smith et al. (1996) or Murray (1996), and each logically rectangular grid cell is divided into four triangles, as illustrated in Fig. 1. All of the thermodynamic and transport variables are given at the center of the cell, velocity is defined at the corners, and the stress tensor is constant across each triangle. We assume contravariant velocity components (velocity components aligned along grid lines). Here, σij may take on four different values within a grid cell. This tends to avoid the grid decoupling problems associated with the B-grid. Note that the rates of strain
The velocity component equations [see (1), (5), (17), or (10), (11)] are coupled through the strain rate
a. Conjugate gradient solution of the viscous–plastic model
Success of the method hinges on symmetry of the iterating and preconditioning matrices; for this reason we lag the terms ±βu during the solution of (18) and (19). This treatment of the Coriolis term, which restricts the time step to about 2 hours for accuracy, might be remedied by applying a predictor–corrector method to these terms as in Zhang and Hibler (1996). This and other improvements to the VP time stepping scheme are reserved for future work.
We have employed a simple linearized Backward-Euler time discretization scheme for (18) and (19). Other methods for dealing with the nonlinearity, such as those employed by Hibler (1979) and Zhang and Hibler (1996), are somewhat more accurate but have their own difficulties. The numerical method of Hibler (1979), which we will refer to as “H79,” iteratively solves the system (10) and (11) at each time step with successive overrelaxation, utilizing a predictor–corrector method to march the equations in time. Specifically, predicted velocities at time level n + 1/2 are used to compute the coefficients of the linearized terms (namely, ζ, η, α, and β) before advancing to the next time level. Hibler and Ackley (1983) found a splitting problem with this procedure in cases of small nonlinear viscosities (free drift), which was corrected by a modified averaging procedure.
As with the predictor–corrector scheme, problems also arise in methods that use numerical spatial splitting and, in particular, in those methods that do not treat the entire strain rate tensor implicitly. For example, Zhang and Hibler (1996) also use successive overrelaxation to solve (10) and (11), along with a predictor–corrector time discretization scheme similar to that of Hibler (1979). In this case, however, the cross derivative terms are treated at time level n instead of n + 1, and the equations decouple. Then the equations for uij are solved iteratively along an entire row (i.e., constant j) before continuing to the next row, and the equations for υij are solved similarly along columns. We will refer to this method as “ZH96.” Stössel et al. (1994) found that treating the diagonal part of the strain rate tensor implicitly and the off-diagonal terms explicitly produced anomalous ice drifts of 6 cm s−1. For the conjugate gradient method described above, the strain rate tensor remains unsplit.
b. The elastic–viscous–plastic model
Discretization in time of the momentum equation (1) is analogous to that of (18) and (19), except that the stress tensor is determined prognostically, and both (1) and (17) are subcycled with an effective EVP time step of length Δte = Δtζ/N for some integer N > 1 and time interval Δtζ. That is, N smaller time steps are taken with (1) and (17), holding η and ζ constant, for each time interval [tn, tn + Δtζ]. Typically, Δtζ = Δtυ, so that Δtυ is often both the viscous–plastic implicit time step and the interval at which viscosity is updated in the EVP model. Subcycling maintains the time scale on which the viscous–plastic material characteristics are changing, ensuring that the VP and EVP formulations are equivalent in the limit Δte → 0, as will be seen later.
4. Heuristic analysis of the elastic–viscous–plastic model
a. Simplified model description
b. Forced response
c. Choosing appropriate parameters
Conditions (35)–(40) may be more easily understood graphically. Figure 3 illustrates the domains of accuracy for the two models, using k2Δx2Δte/Tυ as the ordinate and ωΔte as the abscissa. In general, the EVP domain is larger than the VP domain: if 1/N < Tυ/Te, then the VP domain is entirely contained within the EVP domain.
It is interesting that (42) implies that a suitable subcycling time step Δte is proportional to the harmonic average of Δtυ and the viscous timescale. This result highlights the benefits of the EVP model: the EVP time step may be orders of magnitude larger than the explicit VP time step. Since there is an entire range of viscous timescales associated with the very large range of effective viscosity coefficients, it may be appropriate to choose some intermediate value of the viscous timescale to use in estimating E. The larger the value of E one chooses, the larger the role of the elastic versus the visco–plastic strain rates and the shorter the time step Δte must be.
5. A one-dimensional test problem
a. Steady state
This solution is confirmed by numerical results, to be shown shortly. We now present a series of simulations that explore and compare the behavior of the EVP and VP models. Unless otherwise noted, parameter values for the simulations in this section are those given in Table 3. The 1D solution was obtained with a simple numerical code that integrates (45)–(48) to steady state.
As predicted by our analysis, the steady-state solution is composed of line segments, illustrated in Fig. 4 (labeled “1D”). Figure 4 also presents corresponding numerical solutions of this problem from the 2D models. Due to the imposed land mask, the numerical solutions remain fundamentally two-dimensional, as illustrated in Fig. 5, and therefore not exactly comparable to the 1D solution. Implementing Neumann or periodic boundary conditions in the SOR viscous–plastic codes in order to make the solution more one-dimensional would have been time consuming and not necessary for our purposes. The four 2D models produce remarkably similar steady-state solutions.
The 2D equations were solved on a 40 × 100 grid of square cells (Δx = Δy = 12.7 km), and the cross sections shown are centrally located in the y direction (j = 50). The integration began with a uniform ice field at rest, no-slip conditions were maintained along all four boundaries, and all of the forcing terms were replaced by a single stress τ = (τ, 0). The CPU times shown in Table 4 represent the time used for the dynamics calculation alone; for each case, 31 s were spent in other sections of the calculation and are not included in the table. These calculations were performed by a CRAY Y-MP8/8128 supercomputer. The models were integrated for 2700 simulated hours, taking 450 time steps with Δtζ = Δtυ = 21 600 s. The EVP dynamics were subcycled 72 times for each viscous–plastic time step, thus taking an effective EVP time step of length Δte = 300 s. The EVP numerical model is nearly 40 times more efficient than the original VP code on this test problem.
The elastic and conjugate gradient solutions shown in Figs. 4 and 5 are at steady state; the others are not. Since the corresponding EVP solution is essentially identical to Fig. 5, it is not shown. Doubling the size of the domain from 40 × 100 to 40 × 200, keeping the resolution the same, reduces the magnitude of the EVP steady-state velocity to about 38 cm s−1, closely approximating the 1D numerical solution.
b. Transient behavior
Without subcycling, Δte = Δtζ and the elastic waves do not damp out within the viscous–plastic time step. The EVP results are then quite energetic for larger time steps, as illustrated in Fig. 8. As the time step approaches zero, however, the solutions converge to the reference solution. Furthermore, the two models produce identical results when Δtυ and Δtζ are much shorter than the viscous–plastic stability limit, regardless of subcycling.
Poor adjustment of the VP model has been noticed previously. Hibler (1979) remarks that the viscous–plastic rheology is slow to converge to steady state and requires several time steps with constant forcing to respond accurately. Similarly, Flato and Hibler (1992) note that even the cavitating fluid model should be subcycled several times without changing the forcing. However, many numerical simulations that utilize the viscous–plastic rheology, including numerous sensitivity studies, use 1-day time steps with daily varying winds (e.g., Hibler and Walsh 1982; Hibler and Ackley 1983; Walsh et al. 1985; Ip et al. 1991; Riedlinger and Preller 1991; Chapman et al. 1994). These wind stresses may vary significantly on timescales of a day or so. For example, the wind stress imposed in this example is less than 0.1 dyn cm−2. Since the initial change in wind stress occurs over the first time step (6 h), this is equivalent to a change in the applied wind stress of 0.4 dyn cm−2 per day. The physical wind stress may vary as much as 5 dyn cm−2 per day (Coon 1980), an order of magnitude larger. Not surprisingly, we observe that when integrated with 1-day time steps, the VP numerical model exhibits a weak response to strongly varying winds. The improved transient behavior of the EVP model enhances its ability to capture the response of the ice to such variations in the stress. We will explore the models’ responses to more realistic, time-dependent forcing in the next section.
Both the viscous–plastic transition to steady state and the magnitude of u at steady state depend on ice concentration, as shown in Fig. 9, since the maximum viscosity ζmax varies with compactness as cexp[c*(1 − c)] through the pressure P. Because of this exponential dependence on c, P and ζmax are about two orders of magnitude less for ice concentrations of 0.8 than for 0.9, and therefore the ice rheology is immaterial for c < 0.8, and one cannot distinguish between elastic and viscous–plastic models.
All of the calculations reported here were done with c = 0.9. Holland et al. (1993) point out that shear stress becomes significant for ice concentrations greater than about 0.9. Furthermore, while open water typically exists year round throughout the Arctic, both Arctic and Antarctic ice concentrations are predominantly greater than 90% during the winter (Stössel and Claussen 1993; Gloersen et al. 1992).
6. A two-dimensional problem
The model equations were integrated for 25 simulated days from rest with a time step Δtυ = 1 day, on a 40 × 40 grid of square cells (Δx = Δy = 12.7 km). Such a large time step is not feasible for the EVP dynamics model; for this case, Δtζ = 6 h and Δte = 216 s. These time steps were chosen to illustrate the VP model’s inaccuracy for conditions under which it is often used, and the improvement offered by the EVP formulation. Strictly speaking, results from the various codes are comparable only for very small Δtυ and Δtζ, although we observe in Figs. 10 and 11 that the values of Δtυ and Δtζ used here are sufficiently small to produce comparable results.
These results differ slightly from a time-accurate reference solution, which we define as that produced by the conjugate gradient method with a time step of 60 s. In Fig. 12, we present the differences of domain-averaged kinetic energies per unit mass for each of the methods with that of the reference solution. This comparison indicates that while all of the methods reach a quasi-steady state, the EVP model is much more accurate during the initial “spin up” from rest, and suggests that the EVP model will behave significantly better under the severe wind forcing conditions observed in the polar regions. For example, Arctic winds have been observed to change as much as 350% in a three-day period (Reynolds 1984), and the ice edge may move 35 km day−1 under gale conditions (Roed and O’Brien 1983). In general, geostrophic winds are responsible for 60%–80% of the daily ice variance (Serreze et al. 1989). On these timescales, it is essential that a numerical model for ice dynamics respond accurately to the imposed forcing.
Furthermore, the magnitude of the differences between the viscous–plastic model solutions and the reference case in Fig. 12 indicate that the VP models are slow to respond to more typical forcing variations. The kinetic energy of the H79 solution is better than the conjugate gradient solution by about a factor of 2 due to effectively two iterations of the linearization being taken in the predictor–corrector method used for the time stepping. Incorporating a predictor–corrector method into the time discretization of the conjugate gradient numerical model would improve its accuracy to that of the H79 model, but degrade its efficiency. Regardless, neither VP model is as accurate as the EVP model.
The CPU times given in Table 6 represent the time used for the dynamics calculation alone; the 4 s spent performing I/O for each case is not included in the table. Implementing a two-step time discretization scheme for the conjugate gradient VP numerical model would improve its forcing response to roughly the level of the H79 code and slow it down by approximately a factor of 2. Note that for the Δtυ = 60 s calculation, the conjugate gradient dynamics used 1379 s CPU. We have not made the corresponding calculation with the H79 method, but based on the figures in Table 6, the H79 model would have taken about 12 times longer, or 4.5 h, to perform this calculation. Thus, the standard VP model would require several CPU hours to reach the level of accuracy obtained with Δtυ = 60 s, which the EVP model simulates fairly well in only 41 CPU seconds using Δtζ = 6 h and Δte = 216 s.
7. Summary
Despite its physical and computational problems, the nonlinear viscous–plastic rheology proposed by Hibler (1979) is the most widely accepted model for sea ice dynamics. In the model’s physical description, the ice viscosity suffers a severe singularity: treated as a viscous fluid, rigid sea ice has infinite viscosity. Hibler regularized this problem by setting a maximum viscosity bound, thereby allowing the ice to creep slowly rather than being completely rigid. Even so, the viscosity ranges over many orders of magnitude, and integrating the implicit VP numerical model requires large computational resources, particularly for high resolution grids on parallel architectures. Using smaller maximum viscosity values increases the model’s computational efficiency but produces less accurate results. Our explicit elastic–viscous–plastic model utilizes an elastic mechanism in regions of rigid ice to significantly increase the computational efficiency of the VP numerical model. For comparison purposes, we have chosen to retain the maximum viscosity bound for the results presented here. In this paper we also present a fast, though still implicit, conjugate gradient method for solving the VP equations. Although the conjugate gradient method’s efficiency is comparable to the EVP method’s on serial machines, the explicit EVP model will be substantially more efficient on parallel computers.
Furthermore, due to its semi-implicit treatment of the ice rheology, the standard numerical formulation of the VP model has very poor time response for time steps typically used by researchers in the field, which are often as long as a day. Our investigation of a simplified, one-dimensional version of the VP model indicates that the viscous–plastic model behavior is acceptable only for wind stresses that vary slowly. However, for wind stresses that vary significantly on timescales of a day, the viscous–plastic model response is weak.
This computational pathology may be resolved by improving the numerical method or by changing the physical parameterization in the model. The EVP model represents a combination of these approaches: its (albeit nonphysical) elastic waves enable the use of an efficient, explicit numerical method. We observe improved transient behavior of the solutions, enhancing the model’s ability to capture the ice response to variations in the imposed stress. However, because the EVP model is based on the same linearized viscous–plastic rheology as the VP model, it may inherit similar problems in some parameter regimes.
We have shown that a large range of the elastic wave parameter E exists for which the EVP numerical method is both stable and efficient. In particular, this allows the elastic time step to be orders of magnitude larger than the viscous–plastic timescale in areas of rigid ice. Several considerations must be weighed when choosing the model parameters. The timescale of the external forcing places an upper bound on Δtυ or Δtζ. The choice of the subcycling time step Δte is based on considerations of efficiency and accuracy; some guidelines for choosing Δte are given in section 4. The parameter values used in this paper, namely for Δtζ, E, and Δte, are representive of suitable values that improve both the numerical efficiency and accuracy of the viscous–plastic ice model. A more complete parameter sensitivity study will be reported later.
Other numerical concerns involve maintaining operator symmetry and energy dissipation properties in the discretization of the stress tensor, which arises from a variational principle. Dividing the grid cells into four triangles for spatial discretizations results in higher resolution of the stress tensor and viscosity fields than of thickness and velocity. These numerical improvements, along with the formulation of the EVP model, have resulted in a fast, efficient model of sea ice dynamics well suited to climate studies on parallel machines.
We have coupled the EVP dynamics model to thermodynamic and transport components and will be testing this ice model with daily atmospheric fluxes and validating it with remotely sensed and in situ observations. More complete descriptions of the thermodynamics and transport components and results from the validation of the complete sea ice model are forthcoming.
Acknowledgments
We wish to thank A. Semtner for providing the ZH96 code, M. Maltrud and C. A. Lai for Arctic datasets, and A. Hagberg for many helpful discussions. We are grateful to S. Piacsek for providing Arctic wind stress data and comments on the manuscript. This work has been supported by the DOE CHAMMP Program. We also thank the Center for Nonlinear Studies and the DOE HPCC-NLS program for technical support.
REFERENCES
Ashby, S. F., T. A. Manteufel, and P. E. Saylor, 1990: A taxonomy for conjugate gradient methods. SIAM J. Numer. Anal.,27, 1542–1568.
Bratchie, I., 1984: Rheology of an ice-floe field. Ann. Glaciol.,5, 23–28.
Bryan, K., S. Manabe, and R. L. Pacanowski, 1975: A global ocean–atmosphere climate model. Part II: The oceanic circulation. J. Phys. Oceanogr.,5, 30–46.
Campbell, W. J., 1965: The wind-driven circulation of ice and water in a polar ocean. J. Geophys. Res.,70, 3279–3301.
Chapman, W. L., W. J. Welch, K. P. Bowman, J. Sacks, and J. E. Walsh, 1994: Arctic sea ice variability: Model sensitivities and a multidecadal simulation. J. Geophys. Res.,99, 919–935.
Colony, R., and R. S. Pritchard, 1975: Integration of elastic–plastic constitutive laws. AIDJEX Bull.,30, 55–80.
Coon, M. D., 1980: A review of AIDJEX modeling. Sea Ice Processes and Models, R. S. Pritchard, Ed., University of Washington Press, 12–25.
——, S. A. Maykut, R. S. Pritchard, D. A. Rothrock, and A. S. Thorndike, 1974: Modeling the pack ice as an elastic-plastic material. AIDJEX Bull.,24, 1–105.
Dukowicz, J. K., 1997: Comments on “Stability of the viscous–plastic sea ice rheology.” J. Phys. Oceanogr.,27, 480–481.
——, R. D. Smith, and R. C. Malone, 1993: A reformulation and implementation of the Bryan–Cox–Semtner ocean model on the connection machine. J. Atmos. Oceanic Technol.,10, 195–208.
——, ——, and ——, 1994: Implicit free-surface method for the Bryan–Cox–Semtner ocean model. J. Geophys. Res.,99, 7991–8014.
Elman, H. C., 1994: Iterative methods for linear systems. Large-Scale Matrix Problems and the Numerical Solution of Partial Differential Equations, J. Gilbert and D. Kershaw, Eds., Advances in Numerical Analysis, Vol. III, Clarendon Press, 69–177.
Felzenbaum, A. I., 1961: The theory of steady drift of ice and the calculation of the long period mean drift in the central part of the Arctic Basin. Probl. North,2, 13–44. [Originally in Problemy Severa, No.2, pp. 16–46, 1958.].
Flato, G. M., and W. D. Hibler III, 1989: On a simple sea-ice dynamics model for climate studies. Ann. Glaciol.,14, 72–77.
——, and ——, 1990: A cavitating fluid sea ice model. Cold Reg. Res. Eng. Lab. Monogr., No. 90-1,.
——, and ——, 1992: Modeling pack ice as a cavitating fluid. J. Phys. Oceanogr.,22, 626–651.
Gloersen, P., W. J. Campbell, D. J. Cavalieri, J. C. Comiso, C. L. Parkinson, and H. J. Zwally, 1992: Arctic and Antarctic sea ice, 1978–1987: Satellite passive-microwave observations and analysis. NASA Tech. Rep. SP-511, 290 pp. [Available from Laboratory for Hydrospheric Processes, NASA Goddard Space Flight Center, Greenbelt, MD 20771.].
Gray, J. M. N. T., and L. W. Morland, 1994: A two-dimensional model for the dynamics of sea ice. Philos. Trans. Roy. Soc. London A,347, 219–290.
Hibler, W. D., III, 1974: Differential sea ice drift II: Comparison of mesoscale strain measurements to linear drift theory predictions. J. Glaciol.,13, 457–471.
——, 1979: A dynamic thermodynamic sea ice model. J. Phys. Oceanogr.,9, 817–846.
——, 1980: Modeling a variable thickness sea ice cover. Mon. Wea. Rev.,108, 1943–1973.
——, 1986: Ice dynamics. chap. 9, The Geophysics of Sea Ice, N. Untersteiner, Ed., NATO ASI Series B: Physics, Plenum Press, 577–640.
——, and W. B. Tucker III, 1979: Some results from a linear viscous model of the Arctic ice cover. J. Glaciol.,22, 293–304.
——, and J. E. Walsh, 1982: On modeling seasonal and interannual fluctuation of Arctic sea ice. J. Phys. Oceanogr.,12, 1514–1523.
——, and S. F. Ackley, 1983: Numerical simulation of the Weddell Sea pack ice. J. Geophys. Res.,88, 2873–2887.
——, and K. Bryan, 1987: A diagnostic ice–ocean model. J. Phys. Oceanogr.,17, 987–1015.
Holland, D. M., L. A. Mysak, D. K. Manak, and J. M. Oberhuber, 1993: Sensitivity study of a dynamic thermodynamic sea ice model. J. Geophys. Res.,98, 2561–2586.
Ip, C. F., W. D. Hibler III, and G. M. Flato, 1991: On the effect of rheology on seasonal sea ice simulations. Ann. Glaciol.,15, 17–25.
Manabe, S., K. Bryan, and M. J. Spelman, 1979: A global ocean–atmosphere climate model with seasonal variation for future studies of climate sensitivity. Dyn. Atmos. Oceans,3, 393–426.
Murray, R. J., 1996: Explicit generation of orthogonal grids for ocean models. J. Comput. Phys.,126, 251–273.
Nikiforov, Y. G., A. M. Gukovich, Y. N. Yefimov, and M. A. Romanov, 1967: Principles of a method for calculating the ice redistribution under the influence of the wind during the navigation period in the Arctic seas. AIDJEX Bull.,3, 40–64. [Originally in Tr. Arkt. Antarkt. Inst.,257, pp. 5–25, 1967.].
Oberhuber, J. M., 1993a: Simulation of the Atlantic circulation with a coupled sea ice–mixed layer–isopycnal general circulation model. Part I: Model description. J. Phys. Oceanogr.,23, 808–829.
——, 1993b: Simulation of the Atlantic circulation with a coupled sea ice–mixed layer–isopycnal general circulation model. Part II: Model experiment. J. Phys. Oceanogr.,23, 830–845.
Parkinson, C. L., and W. M. Washington, 1979: A large-scale numerical model of sea ice. J. Geophys. Res.,84, 311–337.
Pritchard, R. S., 1975: An elastic-plastic constitutive law for sea ice. J. Appl. Mech.,42E, 379–384.
——, M. D. Coon, and M. G. McPhee, 1977: Simulation of sea ice dynamics during AIDJEX. J. Pressure Vessel Tech.,99J, 491–497.
Reuss, A., 1930: Berucksichtigung der elastischen formanderung in der plastizitats theorie. Z. Ang. Math. Mech.,10, 266.
Reynolds, M., 1984: On the local meteorology at the marginal ice zone of the Bering Sea. J. Geophys. Res.,89, 6515–6524.
Riedlinger, S. H., and R. H. Preller, 1991: The development of a coupled ice-ocean model for forecasting ice conditions in the Arctic. J. Geophys. Res.,96, 16 955–16 977.
Roed, L. P., and J. J. O’Brien, 1983: A coupled ice-ocean model of upwelling in the marginal ice zone. J. Geophys. Res.,88, 2863–2872.
Rothrock, D. A., 1975a: The steady drift of an incompressible Arctic ice cover. J. Geophys. Res.,80, 387–397.
——, 1975b: The energetics of the plastic deformation of pack ice by ridging. J. Geophys. Res.,80, 4514–4519.
Serreze, M. C., R. G. Barry, and A. S. McLaren, 1989: Seasonal variations in sea ice motion and effects on sea ice concentration in the Canada Basin. J. Geophys. Res.,94, 10 955–10 970.
Smith, R. D., J. K. Dukowicz, and R. C. Malone, 1992: Parallel ocean general circulation modeling. Physica D,60, 38–61.
——, S. Kortas, and B. Meltz, 1997: Curvilinear coordinates for global ocean models. J. Comput. Phys., in press.
Stössel, A., and M. Claussen, 1993: On the momentum forcing of a large-scale sea-ice model. Climate Dyn.,9, 71–80.
——, J. M. Oberhuber, and E. Maier-Reimer, 1994: Sensitivites of the sea-ice component of two ocean general circulation models. Max-Plank-Institut für Meteorologie, Hamburg, Tech. Rep. 138, 44 pp. [Available from Max-Planck-Institute Für Meteorologie, Bundesstrasse 55, 20146 Hamburg, Germany.].
Walsh, J. E., W. D. Hibler III, and B. Ross, 1985: Numerical simulation of Northern Hemisphere sea ice variability, 1951–1980. J. Geophys. Res.,90, 4847–4865.
Warn-Varnas, A., R. Allard, and S. Piacsek, 1991: Synoptic and seasonal variations of the ice-ocean circulation in the Arctic: A numerical study. Ann. Glaciol.,15, 54–62.
Zhang, J., and W. D. Hibler, 1996: On an efficient numerical method for modeling sea ice dynamics. J. Geophys. Res., in press.
APPENDIX A
Numerical Formulations
APPENDIX B
Stability of the 1D Equations
APPENDIX C
VP Model Adjustment to Imposed Forcing
Constants and parameters used in the dynamics equations.
Definitions of other symbols used in the dynamics equations, and their interdependencies.
Initial values and parameters for the tests shown in Fig. 4. The error tolerance on the residual for the VP implicit schemes is given by err.
Estimated total CPU times for the dynamics calculations by each of the four models and the corresponding average CPU time spent for each of the 4000 grid cells for the tests shown in Fig. 4.
Initial values and parameters for the 2D tests.
Estimated total CPU times for the 2D tests by each of the models and the corresponding average CPU time spent for each of the 1600 grid cells.