• Keine Ergebnisse gefunden

Sea ice dynamics solvers in the MITgcm

N/A
N/A
Protected

Academic year: 2022

Aktie "Sea ice dynamics solvers in the MITgcm"

Copied!
28
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

Sea ice dynamics solvers in the MITgcm

Martin Losch (Alfred-Wegener-Institut, Bremerhaven) Jean-Michel Campin (MIT, Cambridge, MA)

COMMODORE, Paris, 2018

(2)

Complicated dynamics

which satellite?

(3)

Complicated dynamics

MITgcm (Menemenlis, Hill)

(4)

Outline: Sea-ice solvers in MITgcm

• Picard solvers (LSR, Krylov)

• JFNK solver

• EVP solvers: mEVP, aEVP

• new MEB rheology in the pipeline

(5)

sea ice dynamics are very non-linear

m ∂u

t = ∇ ⋅ σ + R , R = other terms with σ

ij

= P

2Δ { 2· ϵ

ij

e

−2

+ [ (1 − e

2

)(· ϵ

11

+ · ϵ

22

) − Δ ] δ

ij

} with abbreviations

Δ = (· ϵ

11

+ · ϵ

22

)

2

+ e

−2

[ (· ϵ

11

ϵ ·

22

)

2

+ 4· ϵ

12

] ϵ ·

ij

= 1

2 (

u

i

x

j

+ ∂ u

j

x

i

) (strain rates)

m ∂u

t ∝∂

x

i

( P Δ

u

i

x

j

) + similar terms

(6)

solution techniques: Picard method

• traditional method, e.g., PSOR, Hibler

(1979), LSOR, Zhang and Hibler (1997), (Gauss-Seidel) for linear solver

• Krylov method for linear solver (Lemieux and Tremblay, 2009), requires preconditioner

• stable, but slow

A(u)u = b

⇒ solve A(u n −1 ) ⋅ u n = b

(7)

solution techniques: JFNK solver

• better (quadratic) convergence near minimum (Lemieux et al. 2010, 2012, Losch et al 2014)

• preconditioner for Krylov solver necessary

• expensive

• unstable, especially at high resolution

• stabilization (e.g. Mehlmann and Richter 2017, involves mixing JFNK and Picard methods)

F(u) = A(u)ub F(u

n

) = F(u

n−1

) + F′

un−1

δ u = 0

!

⇒ solve F′

n−1

δ u = − F(u

n−1

) ⇒ u

n

= u

n−1

+ δ u

(8)

Picard vs. JFNK

(9)

Picard vs. JFNK

structures[8].Fig. 4illustrates what we mean by sharp solution structures. It shows the shear deformation field on 7 January 1990 08Z simulated by the JFNK solver when using the 10-km resolution model and a cnl of 0.001. The shear deformation (second strain rate invariant) is given by

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

@u@x !@@yv

" #2

þ"@u@yþ@@xv#2

r

. As in Maslowski and Lipscomb [28], who used a model with about the same spatial resolution (9 km), our model simulates basin scale linear kinematic features that resemble the observed ones [29]. Note that the existence of these strong velocity gradients is physically based (VP rheology) and is not a consequence of residual errors in the velocity field approximate solution.

Note that for the JFNK solver, the computational efficiency and failure rate depend on the chosen value of rest (Eq. (25)) and that some tuning might slightly modify these results. A largerrest tends to increase the computational efficiency and the failure rate.

The lack of convergence (failures) of the JFNK solver and the standard solver is a global convergence issue. When the ini- tial iterate is ‘‘sufficiently close” to the solution, the solvers always converge. The quality of the initial iterate is determined by the time step compared to the forcing time scale and to the level of convergence of the previous time step solution. A 1- month integration at 40-km resolution with a 1-minute time step (44,640 time steps) forcnl ¼ 0:001 shows that both solvers always converge. Unfortunately, the use of such a small time step represents a prohibitive computational approach. We have not investigated what is the maximum time step allowed (between 1 and 30 min at 40-km resolution) for the solvers to con- verge in all cases.

To illustrate the high convergence rate of the JFNK method as opposed to the ones of Stand-cap and Stand-tanh, Fig. 5 shows the residual norm of the nonlinear system of equations as a function of the iteration (Newton iteration or OL iteration) down to a small residual norm ð10!6Þ. This typical result is for 1 January 1990 18Z. The Stand-cap solver needs in this case 2631 OL iterations to reach a residual norm of 10!6 while it takes 24 Newton iterations for JFNK to satisfy the same criterion.

This might suggest that JFNK is more than a 100 times faster than the Stand-cap solver. This is however not the case because one JFNK iteration involves more calculation (in the fast phase) than one OL iteration. JFNK is & 23 times faster than the Stand-cap solver to reach a residual norm of 10!6. Compared to the Stand-tanh solver, JFNK is 6.4 times faster. The required CPU time for JFNK is 2.41 s, 15.49 s for Stand-tanh and 55.07 s for Stand-cap.

Even though the convergence rate of the JFNK solver is high (especially in the fast phase), it is not quadratic because an inexact Newton approach is used. Asymptotic quadratic convergence could be possible but at the expense of very small cðkÞ values [8].

5.2. Discussion about the robustness of the standard and JFNK solvers

Both standard and JFNK solvers show a lack of robustness. Moreover, the failure rate for both solvers increases as the grid is refined. However, the lack of robustness of the solvers might not be so dramatic for practical considerations. First,cnl ¼ 0:2

0 100 200 300 400 500 600 700 800 900 1000

10−6 10−5 10−4 10−3 10−2 10−1 100 101

iteration

residual norm

Stand−cap Stand−tanh JFNK

Fig. 5. Residual norm (N m!2) of the nonlinear system of equations as a function of the OL iteration (or Newton iteration) on 1 January 1990 18Z. The spatial resolution is 40 km.

J.-F. Lemieux et al. / Journal of Computational Physics 229 (2010) 2840–2852 2849

Lemieux et al (2010)

(10)

“Timing” of solvers

(11)

Does it matter?

after nearly 40 years of simulation: average of Oct, 1995

Losch et al. (2014)

(12)

solution method: EVP variants

• Hunke and Dukowicz (1997)

• does not converge (definitely not to VP, Lemieux et al.

2012, Losch and Danilov 2012)

• adding inertial term to momentum equations fixes

convergence (Lemieux et al. 2012, Bouillon et al 2013)

• m(odified)EVP, a(daptive)EVP (Kimmritz et al 2015, 2016, 2017)

σ

ij

= P

2Δ { 2· ϵ

ij

e

−2

+ [ (1 − e

−2

)(· ϵ

11

+ · ϵ

22

) − Δ ] δ

ij

}

⇔ ( 1 E

σ

ij

t + )

Δ e

2

P σ

ij

+ [

Δ(1 − e

2

)

2 P ( σ

11

+ σ

22

) + Δ

2 ] δ

ij

= · ϵ

ij

(13)

solution method: EVP variants

• Hunke and Dukowicz (1997)

• does not converge (definitely not to VP, Lemieux et al.

2012, Losch and Danilov 2012)

• adding inertial term to momentum equations fixes

convergence (Lemieux et al. 2012, Bouillon et al 2013)

• m(odified)EVP, a(daptive)EVP (Kimmritz et al 2015, 2016, 2017)

σ

ij

= P

2Δ { 2· ϵ

ij

e

−2

+ [ (1 − e

−2

)(· ϵ

11

+ · ϵ

22

) − Δ ] δ

ij

}

⇔ ( 1 E

σ

ij

t + )

Δ e

2

P σ

ij

+ [

Δ(1 − e

2

)

2 P ( σ

11

+ σ

22

) + Δ

2 ] δ

ij

= · ϵ

ij

(14)

New EVP equations

based on Lemieux et al. (2012), Bouillon et al. (2013), add “inertial-like” term to momentum equations

now, with

the discretized equations converge to true VP

or explicit, through the EVP formulation (Hunke and Dukowicz, 1997, Hunke

113

and Lipscomb, 2008) where adding a pseudo-elastic term reduces the time step

114

limitations. A discussion of the convergence issues can be found, for instance,

115

in Bouillon et al. (2013), Kimmritz et al. (2015) and is not repeated here.

116

The suggestion by Bouillon et al. (2013) is equivalent, up to details of treating

117

the Coriolis and the ice-ocean drag terms, to formulating the mEVP method as:

118

p+1 p = 1

⇣ (up) p

, (4)

119

up+1 up = 1 ⇣ t

m r · p+1 + t

m Rp+1/2 + un up

. (5)

120 121

In (5), R sums all the terms in the momentum equation except for the rheology

122

and the time derivative, t is the external time step of the sea ice model set by

123

the ocean model, the index n labels the time levels of the model time, and the

124

index p is that of pseudotime (subcycling step number). The Coriolis term in

125

Rp+1/2 is treated implicitly in our B-grid implementation, but is explicit on the

126

C-grid, and the ice-ocean stress term is linearly-implicit (Cdo|uo up|(uo

127

up+1)). The term ij(up) in (4) implies that the stresses are estimated by (2)

128

based on the velocity of iteration p, and ijp is the variable of the pseudotime

129

iteration. The relaxation parameters ↵ and in (4) and (5) are chosen to satisfy

130

stability constraints, see Bouillon et al. (2013), Kimmritz et al. (2015). They

131

replace the terms 2T / te and ( /m)( t/ te), where T is the elastic damping

132

time scale and te the subcycling time step of standard EVP formulation; the

133

parameter was introduced in Lemieux et al. (2012). If (4) and (5) are iterated

134

to convergence, their left hand sides can be set to zero leaving the VP solution:

135

m t

⇣un+1 un

= r · (un+1) + R, (6)

136

with R := limp!1 Rp+1/2 and un+1 := limp!1 up. While one may introduce

137

a convergence criterion to determine the number of iteration steps, historically,

138

the actual number of pseudotime iterations N is selected experimentally to

139

ensure the accuracy needed. The new velocity un+1 at time step n + 1 is

140

estimated at the last pseudotime step p = N. The initial values for p = 1 are

141

taken from the previous time step n.

142

5

or explicit, through the EVP formulation (Hunke and Dukowicz, 1997, Hunke

113

and Lipscomb, 2008) where adding a pseudo-elastic term reduces the time step

114

limitations. A discussion of the convergence issues can be found, for instance,

115

in Bouillon et al. (2013), Kimmritz et al. (2015) and is not repeated here.

116

The suggestion by Bouillon et al. (2013) is equivalent, up to details of treating

117

the Coriolis and the ice-ocean drag terms, to formulating the mEVP method as:

118

p+1 p = 1

⇣ (up) p

, (4)

119

up+1 up = 1 ⇣ t

m r · p+1 + t

m Rp+1/2 + un up

. (5)

120 121

In (5), R sums all the terms in the momentum equation except for the rheology

122

and the time derivative, t is the external time step of the sea ice model set by

123

the ocean model, the index n labels the time levels of the model time, and the

124

index p is that of pseudotime (subcycling step number). The Coriolis term in

125

Rp+1/2 is treated implicitly in our B-grid implementation, but is explicit on the

126

C-grid, and the ice-ocean stress term is linearly-implicit (Cdo|uo up|(uo

127

up+1)). The term ij(up) in (4) implies that the stresses are estimated by (2)

128

based on the velocity of iteration p, and ijp is the variable of the pseudotime

129

iteration. The relaxation parameters ↵ and in (4) and (5) are chosen to satisfy

130

stability constraints, see Bouillon et al. (2013), Kimmritz et al. (2015). They

131

replace the terms 2T / te and ( /m)( t/ te), where T is the elastic damping

132

time scale and te the subcycling time step of standard EVP formulation; the

133

parameter was introduced in Lemieux et al. (2012). If (4) and (5) are iterated

134

to convergence, their left hand sides can be set to zero leaving the VP solution:

135

m t

⇣un+1 un

= r · (un+1) + R, (6)

136

with R := limp!1 Rp+1/2 and un+1 := limp!1 up. While one may introduce

137

a convergence criterion to determine the number of iteration steps, historically,

138

the actual number of pseudotime iterations N is selected experimentally to

139

ensure the accuracy needed. The new velocity un+1 at time step n + 1 is

140

estimated at the last pseudotime step p = N. The initial values for p = 1 are

141

taken from the previous time step n.

142

5

or explicit, through the EVP formulation (Hunke and Dukowicz, 1997, Hunke

113

and Lipscomb, 2008) where adding a pseudo-elastic term reduces the time step

114

limitations. A discussion of the convergence issues can be found, for instance,

115

in Bouillon et al. (2013), Kimmritz et al. (2015) and is not repeated here.

116

The suggestion by Bouillon et al. (2013) is equivalent, up to details of treating

117

the Coriolis and the ice-ocean drag terms, to formulating the mEVP method as:

118

p+1 p

= 1

(u

p

)

p

,

(4)

119

up+1 up

= 1

⇣ t

m r · p+1

+

t

m Rp+1/2

+

un up

.

(5)

120 121

In (5),

R

sums all the terms in the momentum equation except for the rheology

122

and the time derivative,

t

is the external time step of the sea ice model set by

123

the ocean model, the index

n

labels the time levels of the model time, and the

124

index p is that of pseudotime (subcycling step number). The Coriolis term in

125

Rp+1/2

is treated implicitly in our B-grid implementation, but is explicit on the

126

C-grid, and the ice-ocean stress term is linearly-implicit (C

do|uo up|

(u

o

127

up+1

)). The term

ij

(u

p

) in (4) implies that the stresses are estimated by (2)

128

based on the velocity of iteration

p, and ijp

is the variable of the pseudotime

129

iteration. The relaxation parameters

and in (4) and (5) are chosen to satisfy

130

stability constraints, see Bouillon et al. (2013), Kimmritz et al. (2015). They

131

replace the terms 2T / t

e

and (

/m)( t/ te

), where

T

is the elastic damping

132

time scale and

te

the subcycling time step of standard EVP formulation; the

133

parameter

was introduced in Lemieux et al. (2012). If (4) and (5) are iterated

134

to convergence, their left hand sides can be set to zero leaving the VP solution:

135

m t

⇣un+1 un

=

r ·

(u

n+1

) +

R,

(6)

136

with

R

:= lim

p!1 Rp+1/2

and

un+1

:= lim

p!1 up

. While one may introduce

137

a convergence criterion to determine the number of iteration steps, historically,

138

the actual number of pseudotime iterations

N

is selected experimentally to

139

ensure the accuracy needed. The new velocity

un+1

at time step

n

+ 1 is

140

estimated at the last pseudotime step

p

=

N

. The initial values for

p

= 1 are

141

taken from the previous time step

n.

142

5

p+1

= lim

p!1

p

and u

n+1

:= lim

p!1

u

p

(15)

New EVP equations

based on Lemieux et al. (2012), Bouillon et al. (2013), add “inertial-like” term to momentum equations

now, with

the discretized equations converge to true VP

or explicit, through the EVP formulation (Hunke and Dukowicz, 1997, Hunke

113

and Lipscomb, 2008) where adding a pseudo-elastic term reduces the time step

114

limitations. A discussion of the convergence issues can be found, for instance,

115

in Bouillon et al. (2013), Kimmritz et al. (2015) and is not repeated here.

116

The suggestion by Bouillon et al. (2013) is equivalent, up to details of treating

117

the Coriolis and the ice-ocean drag terms, to formulating the mEVP method as:

118

p+1 p = 1

⇣ (up) p

, (4)

119

up+1 up = 1 ⇣ t

m r · p+1 + t

m Rp+1/2 + un up

. (5)

120 121

In (5), R sums all the terms in the momentum equation except for the rheology

122

and the time derivative, t is the external time step of the sea ice model set by

123

the ocean model, the index n labels the time levels of the model time, and the

124

index p is that of pseudotime (subcycling step number). The Coriolis term in

125

Rp+1/2 is treated implicitly in our B-grid implementation, but is explicit on the

126

C-grid, and the ice-ocean stress term is linearly-implicit (Cdo|uo up|(uo

127

up+1)). The term ij(up) in (4) implies that the stresses are estimated by (2)

128

based on the velocity of iteration p, and ijp is the variable of the pseudotime

129

iteration. The relaxation parameters ↵ and in (4) and (5) are chosen to satisfy

130

stability constraints, see Bouillon et al. (2013), Kimmritz et al. (2015). They

131

replace the terms 2T / te and ( /m)( t/ te), where T is the elastic damping

132

time scale and te the subcycling time step of standard EVP formulation; the

133

parameter was introduced in Lemieux et al. (2012). If (4) and (5) are iterated

134

to convergence, their left hand sides can be set to zero leaving the VP solution:

135

m t

⇣un+1 un

= r · (un+1) + R, (6)

136

with R := limp!1 Rp+1/2 and un+1 := limp!1 up. While one may introduce

137

a convergence criterion to determine the number of iteration steps, historically,

138

the actual number of pseudotime iterations N is selected experimentally to

139

ensure the accuracy needed. The new velocity un+1 at time step n + 1 is

140

estimated at the last pseudotime step p = N. The initial values for p = 1 are

141

taken from the previous time step n.

142

5

or explicit, through the EVP formulation (Hunke and Dukowicz, 1997, Hunke

113

and Lipscomb, 2008) where adding a pseudo-elastic term reduces the time step

114

limitations. A discussion of the convergence issues can be found, for instance,

115

in Bouillon et al. (2013), Kimmritz et al. (2015) and is not repeated here.

116

The suggestion by Bouillon et al. (2013) is equivalent, up to details of treating

117

the Coriolis and the ice-ocean drag terms, to formulating the mEVP method as:

118

p+1 p = 1

⇣ (up) p

, (4)

119

up+1 up = 1 ⇣ t

m r · p+1 + t

m Rp+1/2 + un up

. (5)

120 121

In (5), R sums all the terms in the momentum equation except for the rheology

122

and the time derivative, t is the external time step of the sea ice model set by

123

the ocean model, the index n labels the time levels of the model time, and the

124

index p is that of pseudotime (subcycling step number). The Coriolis term in

125

Rp+1/2 is treated implicitly in our B-grid implementation, but is explicit on the

126

C-grid, and the ice-ocean stress term is linearly-implicit (Cdo|uo up|(uo

127

up+1)). The term ij(up) in (4) implies that the stresses are estimated by (2)

128

based on the velocity of iteration p, and ijp is the variable of the pseudotime

129

iteration. The relaxation parameters ↵ and in (4) and (5) are chosen to satisfy

130

stability constraints, see Bouillon et al. (2013), Kimmritz et al. (2015). They

131

replace the terms 2T / te and ( /m)( t/ te), where T is the elastic damping

132

time scale and te the subcycling time step of standard EVP formulation; the

133

parameter was introduced in Lemieux et al. (2012). If (4) and (5) are iterated

134

to convergence, their left hand sides can be set to zero leaving the VP solution:

135

m t

⇣un+1 un

= r · (un+1) + R, (6)

136

with R := limp!1 Rp+1/2 and un+1 := limp!1 up. While one may introduce

137

a convergence criterion to determine the number of iteration steps, historically,

138

the actual number of pseudotime iterations N is selected experimentally to

139

ensure the accuracy needed. The new velocity un+1 at time step n + 1 is

140

estimated at the last pseudotime step p = N. The initial values for p = 1 are

141

taken from the previous time step n.

142

5

or explicit, through the EVP formulation (Hunke and Dukowicz, 1997, Hunke

113

and Lipscomb, 2008) where adding a pseudo-elastic term reduces the time step

114

limitations. A discussion of the convergence issues can be found, for instance,

115

in Bouillon et al. (2013), Kimmritz et al. (2015) and is not repeated here.

116

The suggestion by Bouillon et al. (2013) is equivalent, up to details of treating

117

the Coriolis and the ice-ocean drag terms, to formulating the mEVP method as:

118

p+1 p

= 1

(u

p

)

p

,

(4)

119

up+1 up

= 1

⇣ t

m r · p+1

+

t

m Rp+1/2

+

un up

.

(5)

120 121

In (5),

R

sums all the terms in the momentum equation except for the rheology

122

and the time derivative,

t

is the external time step of the sea ice model set by

123

the ocean model, the index

n

labels the time levels of the model time, and the

124

index p is that of pseudotime (subcycling step number). The Coriolis term in

125

Rp+1/2

is treated implicitly in our B-grid implementation, but is explicit on the

126

C-grid, and the ice-ocean stress term is linearly-implicit (C

do|uo up|

(u

o

127

up+1

)). The term

ij

(u

p

) in (4) implies that the stresses are estimated by (2)

128

based on the velocity of iteration

p, and ijp

is the variable of the pseudotime

129

iteration. The relaxation parameters

and in (4) and (5) are chosen to satisfy

130

stability constraints, see Bouillon et al. (2013), Kimmritz et al. (2015). They

131

replace the terms 2T / t

e

and (

/m)( t/ te

), where

T

is the elastic damping

132

time scale and

te

the subcycling time step of standard EVP formulation; the

133

parameter

was introduced in Lemieux et al. (2012). If (4) and (5) are iterated

134

to convergence, their left hand sides can be set to zero leaving the VP solution:

135

m t

⇣un+1 un

=

r ·

(u

n+1

) +

R,

(6)

136

with

R

:= lim

p!1 Rp+1/2

and

un+1

:= lim

p!1 up

. While one may introduce

137

a convergence criterion to determine the number of iteration steps, historically,

138

the actual number of pseudotime iterations

N

is selected experimentally to

139

ensure the accuracy needed. The new velocity

un+1

at time step

n

+ 1 is

140

estimated at the last pseudotime step

p

=

N

. The initial values for

p

= 1 are

141

taken from the previous time step

n.

142

5

p+1

= lim

p!1

p

and u

n+1

:= lim

p!1

u

p

(16)

New EVP equations

based on Lemieux et al. (2012), Bouillon et al. (2013), add “inertial-like” term to momentum equations

now, with

the discretized equations converge to true VP

or explicit, through the EVP formulation (Hunke and Dukowicz, 1997, Hunke

113

and Lipscomb, 2008) where adding a pseudo-elastic term reduces the time step

114

limitations. A discussion of the convergence issues can be found, for instance,

115

in Bouillon et al. (2013), Kimmritz et al. (2015) and is not repeated here.

116

The suggestion by Bouillon et al. (2013) is equivalent, up to details of treating

117

the Coriolis and the ice-ocean drag terms, to formulating the mEVP method as:

118

p+1 p = 1

⇣ (up) p

, (4)

119

up+1 up = 1 ⇣ t

m r · p+1 + t

m Rp+1/2 + un up

. (5)

120 121

In (5), R sums all the terms in the momentum equation except for the rheology

122

and the time derivative, t is the external time step of the sea ice model set by

123

the ocean model, the index n labels the time levels of the model time, and the

124

index p is that of pseudotime (subcycling step number). The Coriolis term in

125

Rp+1/2 is treated implicitly in our B-grid implementation, but is explicit on the

126

C-grid, and the ice-ocean stress term is linearly-implicit (Cdo|uo up|(uo

127

up+1)). The term ij(up) in (4) implies that the stresses are estimated by (2)

128

based on the velocity of iteration p, and ijp is the variable of the pseudotime

129

iteration. The relaxation parameters ↵ and in (4) and (5) are chosen to satisfy

130

stability constraints, see Bouillon et al. (2013), Kimmritz et al. (2015). They

131

replace the terms 2T / te and ( /m)( t/ te), where T is the elastic damping

132

time scale and te the subcycling time step of standard EVP formulation; the

133

parameter was introduced in Lemieux et al. (2012). If (4) and (5) are iterated

134

to convergence, their left hand sides can be set to zero leaving the VP solution:

135

m t

⇣un+1 un

= r · (un+1) + R, (6)

136

with R := limp!1 Rp+1/2 and un+1 := limp!1 up. While one may introduce

137

a convergence criterion to determine the number of iteration steps, historically,

138

the actual number of pseudotime iterations N is selected experimentally to

139

ensure the accuracy needed. The new velocity un+1 at time step n + 1 is

140

estimated at the last pseudotime step p = N. The initial values for p = 1 are

141

taken from the previous time step n.

142

5

or explicit, through the EVP formulation (Hunke and Dukowicz, 1997, Hunke

113

and Lipscomb, 2008) where adding a pseudo-elastic term reduces the time step

114

limitations. A discussion of the convergence issues can be found, for instance,

115

in Bouillon et al. (2013), Kimmritz et al. (2015) and is not repeated here.

116

The suggestion by Bouillon et al. (2013) is equivalent, up to details of treating

117

the Coriolis and the ice-ocean drag terms, to formulating the mEVP method as:

118

p+1 p = 1

⇣ (up) p

, (4)

119

up+1 up = 1 ⇣ t

m r · p+1 + t

m Rp+1/2 + un up

. (5)

120 121

In (5), R sums all the terms in the momentum equation except for the rheology

122

and the time derivative, t is the external time step of the sea ice model set by

123

the ocean model, the index n labels the time levels of the model time, and the

124

index p is that of pseudotime (subcycling step number). The Coriolis term in

125

Rp+1/2 is treated implicitly in our B-grid implementation, but is explicit on the

126

C-grid, and the ice-ocean stress term is linearly-implicit (Cdo|uo up|(uo

127

up+1)). The term ij(up) in (4) implies that the stresses are estimated by (2)

128

based on the velocity of iteration p, and ijp is the variable of the pseudotime

129

iteration. The relaxation parameters ↵ and in (4) and (5) are chosen to satisfy

130

stability constraints, see Bouillon et al. (2013), Kimmritz et al. (2015). They

131

replace the terms 2T / te and ( /m)( t/ te), where T is the elastic damping

132

time scale and te the subcycling time step of standard EVP formulation; the

133

parameter was introduced in Lemieux et al. (2012). If (4) and (5) are iterated

134

to convergence, their left hand sides can be set to zero leaving the VP solution:

135

m t

⇣un+1 un

= r · (un+1) + R, (6)

136

with R := limp!1 Rp+1/2 and un+1 := limp!1 up. While one may introduce

137

a convergence criterion to determine the number of iteration steps, historically,

138

the actual number of pseudotime iterations N is selected experimentally to

139

ensure the accuracy needed. The new velocity un+1 at time step n + 1 is

140

estimated at the last pseudotime step p = N. The initial values for p = 1 are

141

taken from the previous time step n.

142

5

or explicit, through the EVP formulation (Hunke and Dukowicz, 1997, Hunke

113

and Lipscomb, 2008) where adding a pseudo-elastic term reduces the time step

114

limitations. A discussion of the convergence issues can be found, for instance,

115

in Bouillon et al. (2013), Kimmritz et al. (2015) and is not repeated here.

116

The suggestion by Bouillon et al. (2013) is equivalent, up to details of treating

117

the Coriolis and the ice-ocean drag terms, to formulating the mEVP method as:

118

p+1 p

= 1

(u

p

)

p

,

(4)

119

up+1 up

= 1

⇣ t

m r · p+1

+

t

m Rp+1/2

+

un up

.

(5)

120 121

In (5),

R

sums all the terms in the momentum equation except for the rheology

122

and the time derivative,

t

is the external time step of the sea ice model set by

123

the ocean model, the index

n

labels the time levels of the model time, and the

124

index p is that of pseudotime (subcycling step number). The Coriolis term in

125

Rp+1/2

is treated implicitly in our B-grid implementation, but is explicit on the

126

C-grid, and the ice-ocean stress term is linearly-implicit (C

do|uo up|

(u

o

127

up+1

)). The term

ij

(u

p

) in (4) implies that the stresses are estimated by (2)

128

based on the velocity of iteration

p, and ijp

is the variable of the pseudotime

129

iteration. The relaxation parameters

and in (4) and (5) are chosen to satisfy

130

stability constraints, see Bouillon et al. (2013), Kimmritz et al. (2015). They

131

replace the terms 2T / t

e

and (

/m)( t/ te

), where

T

is the elastic damping

132

time scale and

te

the subcycling time step of standard EVP formulation; the

133

parameter

was introduced in Lemieux et al. (2012). If (4) and (5) are iterated

134

to convergence, their left hand sides can be set to zero leaving the VP solution:

135

m t

⇣un+1 un

=

r ·

(u

n+1

) +

R,

(6)

136

with

R

:= lim

p!1 Rp+1/2

and

un+1

:= lim

p!1 up

. While one may introduce

137

a convergence criterion to determine the number of iteration steps, historically,

138

the actual number of pseudotime iterations

N

is selected experimentally to

139

ensure the accuracy needed. The new velocity

un+1

at time step

n

+ 1 is

140

estimated at the last pseudotime step

p

=

N

. The initial values for

p

= 1 are

141

taken from the previous time step

n.

142

5

p+1

= lim

p!1

p

and u

n+1

:= lim

p!1

u

p

(17)

New EVP equations

New momentum equations

with

from stability analysis (Kimmritz et al, 2015, 2016).

or explicit, through the EVP formulation (Hunke and Dukowicz, 1997, Hunke

113

and Lipscomb, 2008) where adding a pseudo-elastic term reduces the time step

114

limitations. A discussion of the convergence issues can be found, for instance,

115

in Bouillon et al. (2013), Kimmritz et al. (2015) and is not repeated here.

116

The suggestion by Bouillon et al. (2013) is equivalent, up to details of treating

117

the Coriolis and the ice-ocean drag terms, to formulating the mEVP method as:

118

p+1 p = 1

⇣ (up) p

, (4)

119

up+1 up = 1 ⇣ t

m r · p+1 + t

m Rp+1/2 + un up

. (5)

120 121

In (5), R sums all the terms in the momentum equation except for the rheology

122

and the time derivative, t is the external time step of the sea ice model set by

123

the ocean model, the index n labels the time levels of the model time, and the

124

index p is that of pseudotime (subcycling step number). The Coriolis term in

125

Rp+1/2 is treated implicitly in our B-grid implementation, but is explicit on the

126

C-grid, and the ice-ocean stress term is linearly-implicit (Cdo|uo up|(uo

127

up+1)). The term ij(up) in (4) implies that the stresses are estimated by (2)

128

based on the velocity of iteration p, and ijp is the variable of the pseudotime

129

iteration. The relaxation parameters ↵ and in (4) and (5) are chosen to satisfy

130

stability constraints, see Bouillon et al. (2013), Kimmritz et al. (2015). They

131

replace the terms 2T / te and ( /m)( t/ te), where T is the elastic damping

132

time scale and te the subcycling time step of standard EVP formulation; the

133

parameter was introduced in Lemieux et al. (2012). If (4) and (5) are iterated

134

to convergence, their left hand sides can be set to zero leaving the VP solution:

135

m t

⇣un+1 un

= r · (un+1) + R, (6)

136

with R := limp!1 Rp+1/2 and un+1 := limp!1 up. While one may introduce

137

a convergence criterion to determine the number of iteration steps, historically,

138

the actual number of pseudotime iterations N is selected experimentally to

139

ensure the accuracy needed. The new velocity un+1 at time step n + 1 is

140

estimated at the last pseudotime step p = N. The initial values for p = 1 are

141

taken from the previous time step n.

142

5

modified EVP: α, β = constant, order(300) adaptive EVP: α = β = (4γ)

1/2

αβγ = P

( )

2

A

Δ t

m

(18)

Figure 2: fig:alpha1month

The field in the aEVP computation with NEVP = 500 at the end of 31/03/93 (top left) and 30/09/93 (top right). Time series of maximal and root mean square values of

at the last sub-cycling of each month (bottom).

lution and the mEVP (aEVP) scheme with NEV P = 50 and 200 are 8.710 3m

256

(1.510 2m) and 1.210 2m (6.410 3m), while the rms di↵erences are twice

257

as large (indicating that there are some outliers from the mean di↵erences). The

258

di↵erences in the simulated fields (right column in Fig. 3) are large only in

259

the weaker ice zone. The absolute mean di↵erences from the reference solu-

260

tion for the mEVP (aEVP) scheme for NEV P = 50 and 200 are 4.6 10 8s 1

261

(4.3 10 8s 1) and 3.9 10 8s 1 (3.7 10 8s 1). The rms di↵erences from

262

the reference solution in are three times bigger than the mean values. In

263

contrast, in the central Arctic (aice > 0.99) the mean absolute di↵erences of the

264

12

Parameter α (aEVP, N = 500)

Ki mmri tz, L osch , D an ilo v (2 01 7)

(19)

Convergence to VP solution:

JFNK — aEVP difference in ice thickness (m) at 27 km resolution

Figure 3: fig:results_aEVP_0397

Mean di↵erences JFNK-mEVP with = = 300 (rows 1 and 3) and JFNK-aEVP (rows 2 and 4) for NEVP = 50 (rows 1 and 2) and NEV P = 200 (rows 3 and 4) for March 1997 for the ice thickness (left column) and the field (right column).

14

Figure 3: fig:results_aEVP_0397

Mean di↵erences JFNK-mEVP with = = 300 (rows 1 and 3) and JFNK-aEVP (rows 2 and 4) for NEVP = 50 (rows 1 and 2) and NEV P = 200 (rows 3 and 4) for March 1997 for the ice thickness (left column) and the field (right column).

14

50 EVP iterations 200 EVP iterations

Kimmritz et al. (2017)

−−

(20)

MITgcm as a testbed: scalability

speed up

Losch et al. (2014)

(21)

high resolution simulations

(22)
(23)

EVP “convergence” (in FESOM)

Koldunov et al., submitted manuscript, grid resolution ~ 4 km

(24)

EVP “convergence” (in FESOM)

Koldunov et al., submitted manuscript, grid resolution ~ 4 km

(25)

Convergence to VP solution:

ice thickness (m) at 4.5 km grid

spacing

↵ = ⇣ (c⇡)2 Ac

t m

stability

parameter

depends on grid

spacing and local

ice viscosity

(26)

Maxwell Elasto-Brittle rheology

• violation of a Mohr-Coulomb failure criterion determines a damage parameter

• damage parameter affects ice strength, elasticity

• (Girard et al 2011, Rampal et al 2015, Dansereau et al 2016)

• In the pipeline for the MITgcm

1 E

σ

t + 1 λ σ = K : · ϵ

(27)

not sure if I

should show

this

(28)

Summary

• Viscous Plastic rheology:

- Picard solvers with LSR and Krylov solvers - JFNK solver

- many EVP variants, especially stable EVP algorithms (Kimmritz et al. 2015, 2016)

• Maxwell Elasto-Brittle rheology (not quite, yet)

• all in the same code framework (no confounders in comparisons)

• main issues remain, especially at high resolution:

convergence, stability vs. geophysical plausibility vs.

time to solution

Referenzen

ÄHNLICHE DOKUMENTE

Figure 11: RMSD between the approximate solution obtained with the EVP and the reference solution for three different damping time scales. The spatial resolution is 10 km and

In early October 1998, a rapid transformation of internal tidal waves was observed, accompanied by a reduction of the associated energy at all depth levels except for the

Here, we present new biomarker data from surface sediments related to the modern spatial (seasonal) sea-ice variability in the central Arctic Ocean and adjacent marginal seas..

Figure 4.5 shows maps of mean ice drift and thickness in the Arctic and Southern oceans, as derived from two coupled ice–ocean models operated at the Alfred Wegener Institute (North

The atmospheric response to the overall sea-ice changes indi- cates a circulation pattern characterized by a dipole of low pressure over the central Arctic and high pressure over

• Very variable, thicker ice in 2009 than 2008 – Sea Ice Thinning in the central Arctic. • Yes

• Inphase less dependent on ridge conductivity than Quadrature.

In order to relate results obtained during the different time periods, empirical relationships are established between the length of the sea ice season, derived from the