Loss of Hyperbolicity and Ill-posedness of the Viscous–Plastic Sea Ice Rheology in Uniaxial Divergent Flow

J. M. N. T. Gray Institut für Mechanik, Technische Universität Darmstadt, Darmstadt, Germanyand Department of Mathematics, University of Manchester, Manchester, United Kingdom

Search for other papers by J. M. N. T. Gray in
Current site
Google Scholar
PubMed
Close
Full access

Abstract

Local contact interactions between sea ice floes can be modeled on the large scale by treating the pack as a two-dimensional continuum with granular properties. One such model, which has gained prominence, is the viscous plastic constitutive rheology, using an elliptical yield curve and normal flow law. It has been used extensively in ice and coupled ice–ocean studies over the past two decades. It is shown that in uniaxial flow this model reduces to a system of three quasi-linear first-order partial differential equations, which are hyperbolic in convergent flow and have mixed elliptic/hyperbolic behavior in divergence with two imaginary wave speeds. A linear stability analysis shows that the change in type causes the equations to be unstable and ill posed in uniaxial divergence. The root cause is a positive feedback mechanism that becomes stronger and stronger with smaller wavelengths. Numerical computations are used to demonstrate that fingers form and break the ice into discrete blocks. The frequency and growth rate of the fingers increase as the numerical resolution is increased, which implies that the model does not converge to a solution as the grid is refined. Two new models are proposed that are well posed. The first retains the positive feedback mechanism and introduces higher-order derivatives to suppress the unbounded growth rate of the instability. The second eliminates the positive feedback mechanism, and the instability, by repositioning the elliptical yield curve in principal stress space. Numerical simulations show that this model diverges without becoming unstable.

Corresponding author address: Dr. J. M. N. T. Gray, Department of Mathematics, University of Manchester, Oxford Rd., Manchester M13 9PL, United Kingdom.

Abstract

Local contact interactions between sea ice floes can be modeled on the large scale by treating the pack as a two-dimensional continuum with granular properties. One such model, which has gained prominence, is the viscous plastic constitutive rheology, using an elliptical yield curve and normal flow law. It has been used extensively in ice and coupled ice–ocean studies over the past two decades. It is shown that in uniaxial flow this model reduces to a system of three quasi-linear first-order partial differential equations, which are hyperbolic in convergent flow and have mixed elliptic/hyperbolic behavior in divergence with two imaginary wave speeds. A linear stability analysis shows that the change in type causes the equations to be unstable and ill posed in uniaxial divergence. The root cause is a positive feedback mechanism that becomes stronger and stronger with smaller wavelengths. Numerical computations are used to demonstrate that fingers form and break the ice into discrete blocks. The frequency and growth rate of the fingers increase as the numerical resolution is increased, which implies that the model does not converge to a solution as the grid is refined. Two new models are proposed that are well posed. The first retains the positive feedback mechanism and introduces higher-order derivatives to suppress the unbounded growth rate of the instability. The second eliminates the positive feedback mechanism, and the instability, by repositioning the elliptical yield curve in principal stress space. Numerical simulations show that this model diverges without becoming unstable.

Corresponding author address: Dr. J. M. N. T. Gray, Department of Mathematics, University of Manchester, Oxford Rd., Manchester M13 9PL, United Kingdom.

1. Introduction

Sea ice is an important component of the global climate system, providing a dynamic and thermodynamic coupling between the polar oceans and the lower atmosphere. Simple free drift theories to model the dynamics of sea ice were first proposed by Nansen (1902) and Rossby and Montgomery (1935). These were essentially point mass models, which neglected the resistance of the ice floes to compression, with the result that unphysically thick ice was produced in regions of net mean convergence (Semtner 1976). A particularly important feature of the Hibler (1979) sea ice model is that the pack is treated as a two-dimensional continuum with large-scale material properties to represent the net effect of these floe interactions. The similarity of the highly fractured mosaic of ice floes with granular materials suggested that plasticity might be appropriate. A two-dimensional viscous–plastic rheology (Hibler 1979) was therefore developed from an elliptical yield curve (Coon et al. 1974) and associated flow rule. This has since become the classic sea ice model and has been used extensively in ice and coupled ice–ocean studies. In this paper the viscous-plastic sea ice model (Hibler 1979) is shown to change type as it switches from uniaxial convergence to uniaxial divergence, and the implications of this for the linear stability of the system are investigated.

2. Governing equations

The viscous–plastic sea ice model (Hibler 1979) consists of two continuity equations and one vector momentum balance for the ice concentration A, the ice thickness per unit pack area h, and the two-dimensional in-plane velocity v. The ice concentration and thickness continuity equations are
i1520-0485-29-11-2920-e1
where d/dt = ∂/∂t + v ·  is the two-dimensional Lagrangian time derivative, div and are the two-dimensional divergence and gradient operators, and Sh and SA are thermodynamic source terms. These equations are derived from mass balance and kinematic conditions and neglect the effects of sea ice ridging (Gray and Morland 1994; Gray and Killworth 1996), which occurs during convergence. The momentum balance is
i1520-0485-29-11-2920-e3
where m = ρh is the ice mass per unit pack area, ρ is the density of ice, N is the two-dimensional depth-integrated stress tensor, k is the unit vector normal to the plane of motion, f is the Coriolis parameter, τa is the wind stress, τw is a dynamic ocean drag, and H is the sea surface height.
The wind stress and the ocean drag are modeled with quadratic drag laws
i1520-0485-29-11-2920-e4
where ρa, ρw are the densities of air and water, and ca and cw are the quadratic drag coefficients for air and water, respectively. The geostrophic wind, Ua, and water, Uw, velocities are rotated by atmospheric and water turning angles ϕa and ϕw, respectively.
A viscous–plastic rheology is used to describe the large-scale material properties of the sea ice, in which the depth-integrated stress
i1520-0485-29-11-2920-e6
where D = (v + vT)/2 is the two-dimensional strain-rate tensor, 1 is the unit tensor, ζ and μ are bulk and shear viscosities, and P is the ice strength. In the plastic flow regime the viscosities are
i1520-0485-29-11-2920-e7
where e is the ratio of the minor and major axes of the elliptical yield curve. The normal flow rule implies that
η2e−2γ21/2
which is a function of the strain-rate invariants
i1520-0485-29-11-2920-e9
The ice strength P is a function of the sea ice thickness and concentration
PPhCA
where P* = 5 × 103 N m−2 and C = 20 are constants. At low strain rates the plastic flow regime produces arbitrarily large viscosities. To circumvent this a viscous creep state for low strain rates was introduced in which the viscosities attain maximum values ζmax = 2.5 × 108 P and μmax = ζmaxe−2. All the theoretical results presented in this paper assume that the strain rate is sufficiently large to lie outside the viscous creep domain.

3. Uniaxial flow

Let O(xy) be a rectangular Cartesian coordinate system in the two-dimensional sea ice plane and let the velocity have components u and υ along the x and y axes, respectively. The viscous–plastic sea ice model is now reduced to the case of uniaxial flow parallel to the x axis by assuming that all variables are independent of the coordinate y and the cross-flow velocity component υ is identically zero,
yυ
The uniaxial velocity u is therefore independent of y, and the strain rate components reduce to Dxx = η, Dyy = 0, and Dxy = 0, where η = ux. The functions γ2 = η2/4 and Δ = (1 + e−2)1/2|η|. Substituting these into the constitutive relation (6), it follows that in uniaxial plastic flow the depth-integrated stresses
NxxψP,Nxy
where the piecewise linear function
i1520-0485-29-11-2920-e13
and the constants
i1520-0485-29-11-2920-e14

The normal stress is illustrated as a function of η in Fig. 1. For simplicity uniaxial divergence above the strain rate ηm will be referred to in this paper as divergence, uniaxial convergence below the strain rate −ηm will be referred to as convergence, and strain rates in the range −ηmηηm will be referred to as in the viscous creep state. It follows from (13) that ψ is equal to a positive constant ψd in divergent flow and is equal to a negative constant ψc in convergent flow. These two constant states are joined by a linear transition at low strain rates. The sign reversal of the constant ψ between diverging and converging flow is crucial to the analysis that follows.

In uniaxial flow the ice concentration equation (1), the ice thickness equation (2), and the x component of the momentum balance (3) reduce to
i1520-0485-29-11-2920-e16
where Rx is the residual driving force in the x direction. All derivatives of A, h, and u with respect to time and space are indicated by the subscripts t and x, respectively. Following notational convention the spatial derivative of tensor components are indicated by a preceding comma. It follows from (12) that, provided |η| ≥ ηm, the x derivative of the in-plane stress Nxx is
Nxx,xψPAAxPhhx
where PA and Ph are the partial derivatives of the ice strength P with respect to A and h, respectively. All terms involving derivatives of the dependent variables A, h, and u have, therefore, been written on the left-hand side of the equations in (16)–(18). It follows that in uniaxial flow the viscous–plastic model (Hibler 1979) reduces to a quasi-linear first-order system of partial differential equations.
The residual driving force
RxτxaτxwuρghHx
where the superscript x denotes the uniaxial component. This simplified one-dimensional analysis is of direct relevance to practical sea ice problems, as the uniaxial deformation is maintained provided the y component of the momentum balance is satisfied. This requires that the cross-flow components of the wind stress and ocean drag must be in exact balance with the rotational effect of the Coriolis force,
τyamfuτywu
where the superscript y denotes the cross-flow component. Note, that there are no sea surface height gradients as we have assumed all y derivatives are zero.

4. Classification

Quasi-linear systems of partial differential equations can be classified as being of either hyperbolic or elliptic type (Courant and Hilbert 1962). Physically one would expect a time-dependent problem of this form to give rise to a hyperbolic system in which the characteristic curves are well defined. However, it shall be shown here that the type changes as the ice switches from convergence to divergence.

In this paper the standard classification procedure of Courant and Hilbert (1962) for first-order systems in two independent variables, x and t, is used. This entails writing the system in k × k matrix form by gathering the dependent variables into a vector s, that is,
astbsxr,
where a and b are matrices, r is a vector independent of the derivatives of s, and the subscripts t and x denote differentiation with respect to time and space, respectively. The characteristic determinant of the system,
Qbτa
is then calculated. If τ is a real solution of the eigenvalue problem Q = 0 then τ determines a characteristic curve of the system. If all k roots are real and distinct then all the characteristics are defined and the system is called totally hyperbolic. Conversely, if there are no real solutions to the equation Q = 0, then the system is elliptic.
The system of viscous–plastic sea ice dynamics equations (16)–(18) can be written in the matrix form (22) by defining the dependent variable vector s = (A, h, u)T. It follows that the residual vector r = (SA, Sh, Rx)T and the coefficient matrices are
i1520-0485-29-11-2920-e24
The characteristic determinant of the system is
i1520-0485-29-11-2920-e26
After expansion of the determinant the eigenvalue problem, Q = 0 reduces to
uτρhuτ2ψAPAhPh
and the characteristic directions of the system are given by the roots
i1520-0485-29-11-2920-e28
where
i1520-0485-29-11-2920-e29
A strength hardening material is defined as a material whose ice strength increases with both increasing thickness and concentration. That is, the partial derivatives
PAPh
implying that λ is real. For a strength hardening material in convergent uniaxial flow there are three real distinct characteristic directions (28)–(30) since ψ = ψc < 0. The system is therefore totally hyperbolic and relative to a Lagrangian observer, moving with the velocity of the ice, waves propagate away with speed λ(−ψc)1/2. This is the characteristic plastic wave speed of the system.
The viscous–plastic sea ice rheology (Hibler 1979) uses an ice strength (10), which is of the strength hardening type, and
λ2PρCACA
is independent of the ice thickness h. In convergent motions the plastic wave speed is of the order of 10 m s−1 at near 100% concentration and drops rapidly to 20 cm s−1 at 70% concentration. Other ice strength laws are possible. For instance, one might expect that the ice has no strength when the concentration is zero. This would be described by a strength hardening law of the form
PPhAm
with
λ2PρmAm
which is also independent of the ice thickness h. It follows that at zero ice concentration λ = 0 and the characteristics (28)–(30) are equal, implying that the system looses strict hyperbolicity. This may be an important feature at the boundary of an ice pack.

For a strength-hardening material in uniaxial divergent flow there is one real root (28) and a pair of complex conjugate roots (29) and (30) since ψ = ψd > 0. This situation does not satisfy the definition of either totally hyperbolic or elliptic systems. Courant and Hilbert (1962, p. 176) do not formally classify these intermediate types and note that such systems “do not seem to occur in problems of mathematical physics.” Whitham (1974, section 14.2, p. 489) investigated a similar smaller system with two characteristics that are either both real or both imaginary implying either hyperbolic or elliptic type. He notes that “the elliptic case turns out not to be uncommon” although it “leads to ill-posed problems in the wave propagation context” and unstable periodic wave trains. In the viscous-plastic case the system possesses both hyperbolic and elliptic properties in divergence since there is one characteristic direction along each particle trajectory and relative to this Lagrangian framework there is elliptic behavior.

To illuminate this further, the viscous–plastic sea ice model is now linearized about a state of uniform flow u0 with constant ice concentration A0 and thickness h0. Differentiating (18) with respect to time and adding u0 times (18) differentiated with respect to x allows the x derivatives of (16) and (17) to be substituted to obtain a single second-order partial differential equation,
ûttu0ûxtu20ψλ20ûxxrû, ûxût
for small velocity perturbations û, where λ20 = (APA + hPh)0/(ρh0) and the residual r is a function of lower-order derivatives. Quasi-linear second-order partial differential equations of the form,
ttxtxxd,
are classified (e.g., Sneddon 1957; Meister 1996) as
i1520-0485-29-11-2920-e38
For Eq. (36) the discriminant acb2 = ψλ20 and the linearized viscous–plastic model is hyperbolic in convergent flow and elliptic in divergent flow since ψ changes sign. Here the type definition is clear cut, and the ambiguous intermediate type classification in divergent flow with the full quasi-linear first-order system is avoided.

5. Positive feedback, instability, and ill-posedness

Elliptic boundary value problems usually arise in steady-state situations, where the independent variables are spacelike. Here one of the independent variables t is timelike and the problem cannot be solved as a boundary value problem as this would violate causality. Instead one has to try to advance the equations forward in time from prescribed initial conditions, that is, to treat the elliptic system as if it were an initial boundary value problem. For the sake of simplicity, consider what happens when normal modes, û = u* exp(ωtikx), are introduced to the homogeneous (r = 0) linearized second-order sea ice equation (36). On substitution of these wavelike solutions, a dispersion relation of the form
ω2u0ikωu20ψλ20k2
is obtained, with roots
ωu0ikλ0kψ
In convergence Re(ω) = 0 and the equations are neutrally stable. However, in divergence Re(ω) = ±λ0(ψ)1/2k and there is one root with positive real part implying growth in the solution and linear instability (Gray and Killworth 1995). Large deviations from the uniform state are therefore expected. Furthermore, a necessary and sufficient condition for the initial value problem for an equation of second order in the time differentiation to be well posed (Strikwerda 1989) is that there exists a constant q such that
ωq
for all roots ω. In convergence the viscous–plastic model is therefore well posed. However, in divergence there is a root Re(ω+) = λ0(ψ)1/2k, which increases without bound with increasing wavenumber k. The criterion (41) is therefore violated and the system is ill posed.

The loss of hyperbolicity is synonymous with the ill-posedness of the system, as it is the imaginary wave speeds (29) and (30) that cause the growth rate of the linear instability to increase without bound with increasing wavenumber k. The equations can be integrated until the point when the wave speeds become imaginary and then the predictive capability is lost. The subsequent solution can only be computed when additional higher-order terms are introduced into the theory to make it well posed. It is therefore of importance to understand why the equations are ill posed and to decide whether this represents a model of a physical process taking place in real sea ice packs.

The ill-posedness of the system arises because of a subtle positive feedback loop in the coupled system of equations, which becomes stronger and stronger as the wavenumber is increased. Some understanding of the feedback mechanism can be obtained by considering the effect of neglecting the accelerative terms in the momentum balance (18). If the advection uux is neglected, the linearized partial differential equation for û reduces to
ûttu0ûxtψλ20ûxxrû, ûxût
The discriminant acb2 = ψλ20u20/4, which is negative in convergence, and the system is hyperbolic as before. However, in divergence, when ψd is positive, the system is
i1520-0485-29-11-2920-e43
The convective term therefore assists the positive feedback mechanism but is not the root cause. Consider now what happens when the time derivative ut is neglected in (18) instead. The linearized partial differential equation reduces to
u0ûxtu20ψλ20ûxxrû, ûxût
and the discriminant is strictly negative:
acb2u20
It follows that the system is hyperbolic in both convergence and divergence, and the positive feedback mechanism is destroyed. This can also be seen by substituting normal modes into (44) to give
ωu0ψλ20u0ik
and Re(ω) = 0, implying that the instability is removed. In addition, neglecting the stress divergence term in (18) is equivalent to setting ψ = 0 in (36) and (40) and the equations are parabolic, stable, and the feedback is again eliminated.
It is clear that the positive feedback arises from the coupling of the local acceleration to the strength hardening model through the ice concentration and thickness equations. In a divergent motion the acceleration and pressure gradient set up a transient dynamic resistance to the motion, whose strength is proportional to P. Relative to the ice surrounding a weak area
i1520-0485-29-11-2920-e47
The positive feedback loop is therefore complete, and a small weakness of the ice will grow. It is therefore expected, that as the ice diverges, weaker areas will become ever weaker, eventually fragmenting the ice cover into regions of thick ice separated by regions with no ice or very little ice. Indeed as the wavelength of the weakness becomes infinitesimally small, its amplitude will grow infinitely quickly, and this will produce dynamic discontinuities within the ice, which are sensitive to the smallest perturbation.

6. Numerical simulations

Consider a sea ice motion in which there is no freezing or melting, SA = Sh = 0, and there is no net driving force Rx = 0. The system of equations, (16)–(18), in conservative form reduces to
i1520-0485-29-11-2920-e48
where the piecewise linear function ψ is defined in (13) and the ice strength P is defined in (10).
The continuity equations (48) and (49) have a similar structure that allows a useful reduction for special initial conditions. If the initial thickness is proportional to the initial concentration h(x, 0) = h*A(x, 0), where h* is the constant of proportionality, then h(x, t) = h*A(x, t) for all time. With this special initial condition the governing equations reduce to
i1520-0485-29-11-2920-e51
where the constant of proportionality cancels out and the results are therefore valid for arbitrary initial ice thicknesses (providing they are similar).
It is advantageous to nondimensionalize the equations, and the following scalings are introduced:
xLx̃,tLut̃,uuũ,
where u* = (ψdP/ρ)1/2 and tildes are used to indicate nondimensional variables. Dropping the tildes for simplicity, the final conservative system of nondimensional equations is
i1520-0485-29-11-2920-e54
where
i1520-0485-29-11-2920-e56
the constant Ψ = ψc/ψd < 0 and ηm = 0.001.
Initially a finite symmetric sea ice concentration is prescribed in the region |x| ⩽ 1 with an antisymmetric velocity field to ensure that |ux| > ηm. Outside this region there is no ice. The initial conditions are
i1520-0485-29-11-2920-e57
Boundaries lie at x = ±3 at which the boundary conditions are
Atut
These boundary conditions are sufficient to model the flow provided the moving ice edge does not reach x = ±3.

A total variational diminishing Lax–Friedrich monotone upwind scheme for conservation laws (Yee 1989;Toth and Odstrcil 1996) was used to numerically solve the reduced system (54)–(56) subject to the initial value and boundary conditions (57)–(59). This is a high-resolution shock capturing method that has the advantage that it does not require the use of a Riemann solver and can be applied to any system of conservation laws. A good review of the method is presented by Toth and Odstrcil (1996).

The results of the simulation using an equal grid spacing Δx = 10−2 nondimensional units and a time step Δt = 10−4 nondimensional units are shown in Fig. 2 at six time intervals. The initial concentration is ice free, |x| > 1, and then rises rapidly so that there is a large central region in which the ice concentration is almost constant. At t = 0.1, some slight concentration ripples can be seen. Weak areas, where the concentration is less than in the surrounding ice, develop and grow by the positive feedback mechanism (47). This sends a series of fingers through the ice concentration field, with adjacent regions of high and low concentration ice, that grow in amplitude with increasing time. Eventually the fingers penetrate through the ice sheet, fragmenting it into numerous discrete blocks. The linear instability mechanism selects only the divergent part of the wave train since ψ = ψc < 0 implies linear stability in convergence and the perturbation will decay. The fingers therefore only grow downward and are naturally bounded below by the ice-free state, consistent with the global energy statements of Schulkes (1995) and Dukowicz (1997). The linear stability analysis simply shows that there are large deviations from the uniform state, which is demonstrated numerically here.

The numerical grid and time step resolve the solution to a finite resolution and necessarily filter out the high wavenumber and high-frequency response. However, the growth rate of the linear instability becomes stronger with increasing wavenumber (40), and it is expected that, as the spatial resolution and time step are refined, more of these high growth rate modes will be resolved. Figure 3 shows the results of the simulation on a refined grid with Δx = 2.5 × 10−3 and Δt = 5 × 10−5 at the same time intervals as before. Once again the fingers grow and fragment the ice. The fourfold increase in the resolution also increases the number of fingers by approximately a factor of 4, and by t = 0.5 units they have penetrated farther through the ice than before. The small wavelength and high-frequency perturbations introduced by the numerical discretization and rounding error are amplified by the instability and dominate the numerical solution. This has two important consequences. First, numerical results are not unique as they are dependent on the numerical grid and time step size, and second, the symmetry of a symmetric problem, such as this, is destroyed by the asymmetric rounding and discretization error.

7. Rate-dependent models

The numerical simulations show that it is highly undesirable to have a mixed type ill-posed theory, as it looses its predictive power, and a new theory is required. In this section and the next two new well-posed theories are derived using different approaches. The first is a regularization procedure, which introduces higher-order terms to control the ill-posedness, and the second moves the elliptical yield curve in principal stress space so that the positive feedback mechanism is removed. The two approaches raise a basic physical question about the nature of the sea ice pack: Does the positive feedback mechanism (47) represent realistic behavior or not? The answer is unclear at present. The pack is implicitly assumed to be a highly fractured continuum, so the instability cannot represent the break up of floes in tension. However, it might play an important role in the opening of polynyas in coastal regions.

The first theory, presented in this section, retains the positive feedback mechanism (47) by introducing a rate-dependent constitutive relation, instead of rate-independent plasticity, to regularize the problem. Consider the modified constitutive relation
i1520-0485-29-11-2920-e60
with viscosities
i1520-0485-29-11-2920-e61
where ΔM > 0 is constant. This constitutive relation is asymptotically similar to the Hibler (1979) law, given in (6) and (7), for Δ ≫ ΔM. In fact, the two are identical when ΔM = 0. However, the linear stability properties for ΔM > 0 are quite different. To see this (60) is reduced to the case of uniaxial flow in which η = ux, γ2 = η2/4, Δ = E|η|, and E = (1 + e−2)1/2. The in-plane stress
Nrxxr
where
i1520-0485-29-11-2920-e63
and the positive constant ηM = ΔM/E. In contrast to ψ, defined in (13), ψr is a continuously varying function of η, as shown in the dash–dot graph in Fig. 1. Its derivative with respect to η is
i1520-0485-29-11-2920-e64
and this therefore contributes to the stress gradient
Nrxx,xPxψrrηuxx
If Nrxx,x is substituted into the momentum balance (18), a nonlinear diffusion term is introduced, which will tend to smooth discontinuities away.
As |η| becomes large, relative to ηM, the diffusion coefficient becomes small and we may linearize the system around a state with a small diffusion coefficient ε0 = P0(ψrη)0/(ρh0), which is strictly positive. Combining the linearized equations as before yields the third-order equation
ûttu0ûxtu20ψr0λ20)ûxx0ûxxtu0ûxxxrû,ûxût
where the small parameter ε0 multiplies the highest derivative terms, and we expect singular behavior. Introducing normal modes the dispersion relation
i1520-0485-29-11-2920-e67
which has a root with a positive real part in divergence
i1520-0485-29-11-2920-e68
which is illustrated in Fig. 4. The system is therefore linearly unstable. However, there is an upper bound,
ω+ωmψrd0λ200
and the system is well posed by (41). Note that, if ε0 → 0, the system tends to a rate-independent state and the upper bound ωm → ∞, implying ill-posedness, which is consistent with the previous results. Nonlinear diffusion therefore introduces smoothing into the positive feedback mechanism that suppresses the ill-posedness but does not prevent the linear instability from growing. It follows that the ice will still fragment, but the formation of dynamic discontinuities is prevented.
Introducing rate-dependence into the constitutive relation does not always remove the ill-posedness. Consider the alternative function
i1520-0485-29-11-2920-e70
which is illustrated in Fig. 1 with the dashed line. Its derivative,
i1520-0485-29-11-2920-e71
is negative for |η| large enough. It follows that this model leads to a diffusion coefficient ε0 = P0(ψRdη)0/(ρh0), which is negative, implying antidiffusion. The growth rate (68) is unbounded in both convergence and divergence and the system is always ill posed, as illustrated in Fig. 4.

8. Shift models

The rate-dependent model (60) introduced higher-order derivatives to control the ill-posedness, but the positive feedback mechanism (47) was retained implying that the linear instability still occurs. As it is unclear whether or not this instability actually represents physical behavior in real sea ice packs, it is of interest to see how it can be eliminated completely from the theory.

The root cause of the positive feedback is that tensile stresses develop during divergent motions. A simple translation of the yield curve in principal stress space is sufficient to ensure that the tensile stress, positive feedback, and ill-posedness are eliminated without including higher-order derivatives. Consider the modified constitutive relation
i1520-0485-29-11-2920-e72
where s is the shift parameter. This is a generalization of the constitutive relation (6). The position of the center of the elliptical yield curve is simply moved along the line of equal principal stresses, N1 = N2, by changing the value of the shift parameter as shown in Fig. 5. The Hibler (1979) model is recovered by setting s = 1. In uniaxial flow the in-plane stress
NsxxψsP,
where the piecewise linear function
i1520-0485-29-11-2920-e74
and the constants are
i1520-0485-29-11-2920-e75
The classification procedure proceeds in exactly the same way as before, and the characteristic directions of the system are again given by (28)–(30). Assuming that a strength-hardening law is used, the characteristics are real provided ψs is negative. It follows that in convergence
i1520-0485-29-11-2920-e77
and in divergence
i1520-0485-29-11-2920-e80
Hyperbolicity is therefore guaranteed in both uniaxial convergence and uniaxial divergence provided
sE.
It also follows from the dispersion relation (40) that the system is linearly stable and well posed since Re(ω) = 0 for ψs ⩽ 0. The acceleration and strength-hardening relation now couple to produce a dynamic acceleration with a strength proportional to P. It follows that in a region of slightly weaker ice
i1520-0485-29-11-2920-e84
and the system equilibrates itself. That is, there is a negative feedback loop instead of the positive one (47). A repositioning of the yield curve to avoid the tensile stress regime is sufficient to eliminate the positive feedback mechanism and hence the linear instability.

To demonstrate the behavior of the shifted ellipse model numerical computations for the case, s = 2E + 1, are presented using the nondimensional variables introduced in (53). This implies that ψ̃sd = −1 in this case. The results for Δx = 10−2 and Δt = 10−4 nondimensional units are shown in Fig. 6 at the same time steps as before. The ice now spreads out laterally, driven by the initial velocity and its own internal pressure. The ice concentration decreases slowly in response to the divergent velocity field. No instabilities or fingers are generated, the symmetry is preserved, and the solution does not change with further grid refinement.

9. Conclusions

In uniaxial flow the viscous–plastic sea ice model (Hibler 1979) reduces to a quasi-linear system of three first-order partial differential equations, which are hyperbolic in convergence and have two imaginary wave speeds in divergence, implying mixed elliptic/hyperbolic properties. The mixed problem cannot be solved as an elliptic boundary value problem as this would violate causality. Instead, it has to be treated as an initial boundary value problem, which has wavelike solutions. These are linearly unstable (Gray and Killworth 1995) and, since the growth rate of a small disturbance increases without bound with increasing wavenumber, the equations are ill posed.

The source of the problem is a subtle unchecked positive feedback loop (47) between the acceleration terms and the strength-hardening law, which sets up a transient dynamic resistance to a divergent motion. Areas of weak ice have less transient resistance and diverge more than the surrounding ice, which further reduces their strength, and the weakness grows. As the wavelength of this instability tends to zero, its growth rate tends to infinity, implying that the problem is ill posed. Numerical simulations show that fingers are generated by the instability, which propagate through the ice and eventually fragment it. The numerical results are not unique. As the grid size and time step are refined, the smallest wavelength and highest-frequency responses are amplified by the instability and dominate the numerical solution. That is, the fingers have a shorter wavelength and grow faster as the resolution increases as shown in Figs. 2 and 3.

This demonstrates that it is highly undesirable to have a mixed type ill-posed theory as the systems predictive capability is lost. A new well-posed theory is therefore required, and two new models have been proposed. The first model retains the positive feedback mechanism (47) by introducing rate-dependent higher-order derivatives to regularize the theory and suppress the unbounded grow rate of the instability. The second eliminates the positive feedback loop, and hence the instability, by moving the elliptical yield curve into the third quadrant of principal stress space. This theory is totally hyperbolic and a numerical simulation demonstrates that it remains stable as the ice diverges (see Fig. 6). These two approaches raise a fundamental question as to the physical properties of real sea ice packs: Does the positive feedback mechanism (47) exist or not, and, if so what physical process does it represent? The answer is unclear at present.

Many sea ice simulations use the Hibler (1979) code to produce numerical results. Usually the equations are uncoupled from one another and solved sequentially with updates once a day. Such long time steps are sufficient for the ice to be very close to steady state and the transient resistance, and hence the positive feedback loop, in divergence, are not resolved. The results that are generated are not solutions of the underlying system of partial differential equations that claim to be solved since the high-frequency and small wavelength signals, which have a large effect, are filtered out. This paper demonstrates that as both grid sizes and time steps are reduced in size, the numerical results do not converge to a solution.

Acknowledgments

This research was supported by the Deutsche Forschungsgemeinschaft through the SFB 298 project “Deformation und Versagen bei metallischen und granularen Strukturen.”

REFERENCES

  • Coon, M. D., G. 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.

  • Courant, R., and D. Hilbert, 1962: Methods of Mathematical Physics. Part II. Partial Differential Equations. Interscience Publishers, John Wiley and Sons, 830 pp.

  • Dukowicz, J. K., 1997: Comments on “Stability of the viscous–plastic sea ice rheology.” J. Phys. Oceanogr.,27, 480–481.

  • 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.

  • ——, and P. D. Killworth, 1995: Stability of the viscous–plastic sea ice rheology. J. Phys. Oceanogr.,25, 971–978.

  • ——, and ——, 1996: Sea ice ridging schemes. J. Phys. Oceanogr.,26, 2420–2428.

  • Hibler, W. D., 1979: A dynamic thermodynamic sea ice model. J. Phys. Oceanogr.,9, 815–845.

  • Meister, E., 1996: Partielle Differential-Gleichungen. Eine Einführung für Physiker und Ingenieure in Die Klassische Theorie. Akademie Verlag, 293 pp.

  • Nansen, F., 1902: The oceanography of the north polar basin. The Norwegian polar expedition 1893–1896. Scientific results, Sci. Res.,3, 357–386.

  • Rossby, C. G., and R. B. Montgomery, 1935: The layer of frictional influence in wind and water currents. Papers Phys. Oceanogr., MIT Woods Hole Oceanographic Institution, 100 pp.

  • Schulkes, R. M. S. M., 1995: Asymptotic stability of the viscous–plastic sea ice rheology. J. Phys. Oceanogr.,26, 279–283.

  • Semtner, Jr., A. J., 1976: Numerical simulations of the Arctic Ocean circulation. J. Phys. Oceanogr.,6, 409–425.

  • Sneddon, I. N., 1957: Elements of Partial Differential Equations. McGraw-Hill, 327 pp.

  • Strikwerda, J. C., 1989: Finite Difference Schemes and Partial Differential Equations. Wadsworth & Brooks/Cole, Mathematics Series, 386 pp.

  • Toth, G., and D. Odstrcil, 1996: Comparison of some flux corrected transport and total variational schemes for hydrodynamic and magnetohydrodynamic problems. J. Comput. Phys.,128, 82.

  • Whitham, G. B., 1974: Linear and Nonlinear Waves. J. Wiley & Sons, 636 pp.

  • Yee, H. C., 1989: A class of high resolution explicit and implicit shock capturing methods. NASA TM-101088.

Fig. 1.
Fig. 1.

The in-plane stress Nxx is plotted as a function of the divergence η for uniaxial flow. The thick solid line shows the rate-independent plastic behavior in convergence and divergence with a linear transition between the two states for small strain rates. The dot–dash line shows a monotone increasing regularization of the theory, and the dashed line show a regularization that leads to antidiffusive behavior.

Citation: Journal of Physical Oceanography 29, 11; 10.1175/1520-0485(1999)029<2920:LOHAIP>2.0.CO;2

Fig. 2.
Fig. 2.

The ice concentration A is plotted as a function of space x at a series of time intervals for the nondimensional ill-posed mixed equations, using a grid spacing Δx = 10−2 and a time step of Δt = 10−4 nondimensional units.

Citation: Journal of Physical Oceanography 29, 11; 10.1175/1520-0485(1999)029<2920:LOHAIP>2.0.CO;2

Fig. 3.
Fig. 3.

The ice concentration A is plotted as a function of space x at a series of time intervals for the nondimensional ill-posed mixed equations, using a grid spacing Δx = 2.5 × 10−3 and a time step of Δt = 5 × 10−5 nondimensional units.

Citation: Journal of Physical Oceanography 29, 11; 10.1175/1520-0485(1999)029<2920:LOHAIP>2.0.CO;2

Fig. 4.
Fig. 4.

The growth rate Re(ω) is plotted as a function of the wavenumber k for divergent uniaxial flow. The solid line is for the perfectly plastic model, the dot–dash line for the monotone increasing regularization, and the dashed line for the antidiffusive model.

Citation: Journal of Physical Oceanography 29, 11; 10.1175/1520-0485(1999)029<2920:LOHAIP>2.0.CO;2

Fig. 5.
Fig. 5.

Three yield curves for the modified constitutive relation (72) are plotted in principal stress space.

Citation: Journal of Physical Oceanography 29, 11; 10.1175/1520-0485(1999)029<2920:LOHAIP>2.0.CO;2

Fig. 6.
Fig. 6.

The ice concentration A is plotted as a function of space x at a series of time intervals for the nondimensional hyperbolic shifted ellipse model with ψ̃sd = −1, using a grid spacing Δx = 10−2 and a time step of Δt = 10−4 nondimensional units.

Citation: Journal of Physical Oceanography 29, 11; 10.1175/1520-0485(1999)029<2920:LOHAIP>2.0.CO;2

Save
  • Coon, M. D., G. 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.

  • Courant, R., and D. Hilbert, 1962: Methods of Mathematical Physics. Part II. Partial Differential Equations. Interscience Publishers, John Wiley and Sons, 830 pp.

  • Dukowicz, J. K., 1997: Comments on “Stability of the viscous–plastic sea ice rheology.” J. Phys. Oceanogr.,27, 480–481.

  • 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.

  • ——, and P. D. Killworth, 1995: Stability of the viscous–plastic sea ice rheology. J. Phys. Oceanogr.,25, 971–978.

  • ——, and ——, 1996: Sea ice ridging schemes. J. Phys. Oceanogr.,26, 2420–2428.

  • Hibler, W. D., 1979: A dynamic thermodynamic sea ice model. J. Phys. Oceanogr.,9, 815–845.

  • Meister, E., 1996: Partielle Differential-Gleichungen. Eine Einführung für Physiker und Ingenieure in Die Klassische Theorie. Akademie Verlag, 293 pp.

  • Nansen, F., 1902: The oceanography of the north polar basin. The Norwegian polar expedition 1893–1896. Scientific results, Sci. Res.,3, 357–386.

  • Rossby, C. G., and R. B. Montgomery, 1935: The layer of frictional influence in wind and water currents. Papers Phys. Oceanogr., MIT Woods Hole Oceanographic Institution, 100 pp.

  • Schulkes, R. M. S. M., 1995: Asymptotic stability of the viscous–plastic sea ice rheology. J. Phys. Oceanogr.,26, 279–283.

  • Semtner, Jr., A. J., 1976: Numerical simulations of the Arctic Ocean circulation. J. Phys. Oceanogr.,6, 409–425.

  • Sneddon, I. N., 1957: Elements of Partial Differential Equations. McGraw-Hill, 327 pp.

  • Strikwerda, J. C., 1989: Finite Difference Schemes and Partial Differential Equations. Wadsworth & Brooks/Cole, Mathematics Series, 386 pp.

  • Toth, G., and D. Odstrcil, 1996: Comparison of some flux corrected transport and total variational schemes for hydrodynamic and magnetohydrodynamic problems. J. Comput. Phys.,128, 82.

  • Whitham, G. B., 1974: Linear and Nonlinear Waves. J. Wiley & Sons, 636 pp.

  • Yee, H. C., 1989: A class of high resolution explicit and implicit shock capturing methods. NASA TM-101088.

  • Fig. 1.

    The in-plane stress Nxx is plotted as a function of the divergence η for uniaxial flow. The thick solid line shows the rate-independent plastic behavior in convergence and divergence with a linear transition between the two states for small strain rates. The dot–dash line shows a monotone increasing regularization of the theory, and the dashed line show a regularization that leads to antidiffusive behavior.

  • Fig. 2.

    The ice concentration A is plotted as a function of space x at a series of time intervals for the nondimensional ill-posed mixed equations, using a grid spacing Δx = 10−2 and a time step of Δt = 10−4 nondimensional units.

  • Fig. 3.

    The ice concentration A is plotted as a function of space x at a series of time intervals for the nondimensional ill-posed mixed equations, using a grid spacing Δx = 2.5 × 10−3 and a time step of Δt = 5 × 10−5 nondimensional units.

  • Fig. 4.

    The growth rate Re(ω) is plotted as a function of the wavenumber k for divergent uniaxial flow. The solid line is for the perfectly plastic model, the dot–dash line for the monotone increasing regularization, and the dashed line for the antidiffusive model.

  • Fig. 5.

    Three yield curves for the modified constitutive relation (72) are plotted in principal stress space.

  • Fig. 6.

    The ice concentration A is plotted as a function of space x at a series of time intervals for the nondimensional hyperbolic shifted ellipse model with ψ̃sd = −1, using a grid spacing Δx = 10−2 and a time step of Δt = 10−4 nondimensional units.

All Time Past Year Past 30 Days
Abstract Views 0 0 0
Full Text Views 351 113 36
PDF Downloads 156 50 4