Tadmor Acta2003 Entropy Conservative Schemes PDF
Tadmor Acta2003 Entropy Conservative Schemes PDF
∗
Research was supported by NSF grants DMS01-07917 and DMS01-07428 and by ONR
grant N00014-91-J-1076.
452 E. Tadmor
We extend these results to fully discrete schemes. Here, spatial entropy dissi-
pation is balanced by the entropy production due to time discretization with
a sufficiently small time-step, satisfying a suitable CFL condition. Finally,
we revisit the question of entropy stability for fully discrete schemes using a
different approach based on homotopy arguments. We prove entropy stability
under optimal CFL conditions.
CONTENTS
1 Introduction 452
2 The entropy variables 456
3 Entropy-conservative and entropy-stable schemes 463
4 The scalar problem 464
5 Systems of conservation laws 470
6 Entropy-conservative schemes revisited 482
7 Entropy stability of fully discrete schemes 488
8 Entropy stability by the homotopy approach 494
9 Higher-order extensions 500
References 502
Appendix: Entropy stability of Roe-type schemes 507
1. Introduction
We discuss the stability of difference approximations to conservation laws
and related time-dependent problems. The related problems we have in
mind are governed by additional dissipative and dispersive forcing terms.
Our main focus, however, is devoted to nonlinear convection governed by
hyperbolic systems of conservation laws. In the linear hyperbolic framework,
L2-stability is sought as a discrete analogue for the a priori energy estimates
available in the differential set-up, e.g., Richtmyer and Morton (1967) and
Gustafsson, Kreiss and Oliger (1995); consult the recent Acta Numerica
review by Kreiss and Lorenz (1998). In the present context of nonlinear
problems dominated by nonlinear convection, we seek entropy stability as a
discrete analogue for the corresponding statement in the differential set-up.
The prototype one-dimensional problem consists of systems of conservation
laws, ut +f (u)x = 0. A distinctive feature of this problem is the spontaneous
formation of shock discontinuities. The entropy condition plays a decisive
role in the theory and numerics of such problems (Lax 1972, Smoller 1983,
Dafermos 2000). It requires u to satisfy the additional inequality, U (u)t +
F (u)x ≤ 0, for all admissibleR entropy pairs (U (u), F (u)). It follows that
the total amount of entropy, U (u(·, t) dx, does not increase in time. This
is a generalization of the (weighted) L2 -energy bound encountered in the
Entropy stability 453
linear case. The possibility of strict inequality reflects entropy decay due to
concentration along shock discontinuities.
We consider difference approximations of the general conservative form,
d fν+ 1 − fν− 1
2 2
uν (t) + = 0.
dt ∆xν
Here uν (t) is the numerical solution computed at discrete grid lines (xν , t),
and fν+ 1 ∼ f is a numerical flux based on a stencil of neighbouring grid
2
values, uν−p+1 , . . . , uν+p . We enquire when such schemes are entropy-stable
in the sense of satisfying the corresponding discrete entropy inequality,
d Fν+ 1 − Fν− 1
2 2
U (uν (t)) + ≤ 0.
dt ∆xν
So far we have specified semi-discrete schemes based on spatial differenc-
ing. We will address the question of entropy stability for the semi-discrete
as well as the fully discrete case, taking into account additional temporal
discretization. The extension to the multidimensional set-up and a host of
related problems with additional dissipative and dispersive terms can be
handled in a straightforward manner.
We distinguish between three main tools of the trade in the analysis of
entropy stability: comparison arguments, a homotopy approach and ki-
netic formulations. We will discuss the first two and refer the reader to
Bouchut (2002), Makridakis and Perthame (2003) and the references therein
for recent contributions regarding the third. Most of our discussion will
be devoted to the main approach, based on a comparison principle: we
compare the amount of entropy dissipation produced by a given scheme
against a properly chosen entropy-stable reference. The entropy stability
of solutions to monotone schemes, for example, Harten, Hyman and Lax
(1976), is carried out by a comparison with the (entropy-stable) constant
solution (Crandall and Majda 1980). The class of entropy-stable E-schemes
(Osher 1984) is characterized by having more numerical viscosity than the
entropy-stable Godunov scheme (Tadmor 1984b). And we mention in pass-
ing the kinetic approach presented in Makridakis and Perthame (2003),
which is based on comparison of the corresponding pseudo-Maxwellians.
In Tadmor (1987b), the question of entropy stability was addressed by the
construction of certain entropy-conservative schemes, interesting for their
own sake. We begin, in Section 3, with the construction of these entropy-
conservative schemes. There are two main ingredients: (i) the use of entropy
variables, outlined in Section 2, and (ii) the choice of certain paths of inte-
gration in phase space of these entropy variables. In the scalar case, the nu-
merical fluxes are path-independent, and entropy-conservative schemes are
unique (for a given entropy pair). In Section 4 we study a host of instruc-
tive scalar examples whose entropy stability is verified by comparison with
454 E. Tadmor
There are three prototype examples. In the fully implicit case where we
1
set un+ 2 := un+1 , additional entropy dissipation is introduced by the time
discretization and hence this implicit backward Euler scheme is entropy-
stable whenever the semi-discrete scheme is. In the case of Crank–Nicolson
time discretization, a proper (possibly nonlinear) choice of intermediate val-
1
ues un+ 2 inherits the same unconditional entropy stability properties of
the semi-discrete problem associated with the numerical flux fν+ 1 ; finally,
2
1
the fully explicit case, un+ 2 := un , yields entropy production which needs
to be balanced by entropy dissipation on the spatial part. This balance
is achieved for a mesh ratio satisfying a suitable Courant–Friedrichs–Lewy
∆t
(CFL) condition, ∆x kfu k ≤ Const.
In Section 8 we revisit the question of entropy stability using a completely
different approach, based on homotopy arguments. The results apply to
semi- and fully discrete approximations of scalar and systems of conservation
laws. We prove the entropy stability for a large class of first-order schemes,
this time under an optimal CFL condition. For second-order scalar exten-
sions we refer to Nessyahu and Tadmor (1990, Appendix). The homotopy
argument was introduced by Lax (1971) in the context of the Lax–Friedrichs
scheme.
The entropy stability study is based on comparison with entropy-con-
servative schemes. The entropy-stable schemes discussed so far were lim-
ited by the use of second-order accurate entropy-conservative schemes as a
Entropy stability 455
procedure, which respects both strong and weak solutions of (2.1), is to sym-
metrize ‘on the right’, where (2.2) is replaced by the equivalent statement
>
A(Uuu)−1 = A(Uuu)−1 . (2.3)
To this end, Mock (1980) (see also Godunov (1961)) suggested the follow-
ing procedure. Define the entropy variables
v ≡ v(u) := ∇u U (u). (2.4)
Thanks to the convexity of U (u), the mapping u → v is one-to-one and
hence we can make the change of variables u = u(v), which puts the system
(2.1) into its equivalent symmetric form
∂ ∂
u(v) + g(v) = 0, g(v) := f (u(v)). (2.5)
∂t ∂x
Here, u(·) and g(·) become the temporal and spatial fluxes in the indepen-
dent entropy variables, v, and the system (2.5) is symmetric in the sense
that the Jacobians of these fluxes are, namely
H(v) := uv (v) = H > (v) > 0 and B(v) := gv (v) = B > (v). (2.6)
Indeed, (2.2) holds if and only if there exists an entropy flux function, F =
F (u), such that the following compatibility relation holds:
Uu> fu = Fu> . (2.7)
Consequently, we have
u(v) = ∇v φ(v), φ(v) := hv, u(v)i − U (u(v)) (2.8)
g(v) = ∇v ψ(v), ψ(v) := hv, g(v)i − F (u(v)), (2.9)
where h·, ·i denotes the usual Euclidean inner product. Hence the Jacobians
H(v) and B(v) in (2.6) are symmetric, being the Hessians of φ(v) and ψ(v).
The latter, so-called potential functions, φ(v) and ψ(v), are significant tools
in our discussion below. Observe that the symmetry of B = AH amounts
to the symmetrization ‘on the right’ indicated in (2.3).
Entropy functions play an important role in the stability theory of PDEs
dominated by the nonlinear convection of the type (2.1). We provide be-
low a brief overview and refer the reader to a detailed account in Volpert
(1967), Kružkov (1970), Friedrichs and Lax (1971), Lax (1972), Tartar
(1975), DiPerna (1983), Smoller (1983), Majda (1984), Serre (1999), Dafer-
mos (2000) and LeFloch (2002). We first recall that ‘physically relevant’ so-
lutions of (2.1), are those arising as vanishing viscosity limits, u = lim↓0 u ,
where
ut + f (u )x = (P ux )x. (2.10)
Here P = P (u, ux ) is any admissible viscosity matrix which is H-symmetric
458 E. Tadmor
(at least one) entropy pair. The canonical example is of course the following
one.
These equations govern inviscid polytropic gas dynamics, asserting the con-
servation of the density ρ, the momentum m, and the total energy E.
Here q and p are, respectively, the velocity q := m
ρ and the pressure p =
m2
(γ − 1) · E − 2ρ (where γ is the adiabatic exponent). Harten has shown
that this system of equations is equipped with a family of entropy pairs,
(U, F ). These pairs take the form
U (u) = −ρh(S), F (u) = −mh(S) (2.16)
(Harten 1983b; consult also Tadmor (1986a)). Here S stands for the non-
dimensional specific entropy
S = `n(pρ−γ ), (2.17)
and h = h(S) is any scalar function satisfying
h0 − γh00 > 0, h0 > 0, (2.18)
so that the requirement for U (u) to be convex is met (Harten 1983b). The
corresponding entropy variables are given by
v1 0
p
E + γ−1 ( hh(S)
0 (S) − γ − 1)
h (S)
v ≡ v2 = (1 − γ) · · −m , (2.19)
v3 p
ρ
(1) The Euler equations (2.15) provide us with an example which shows
how the ‘richness’ of the entropy pairs can be used for a stability state-
ment: using the one-parameter family2 (−ρ(S − c)− , −m(S − c)− ),
which is admissible by (2.18), we obtain a minimum entropy principle,
S(x, t) ≥ miny S(y, 0) (consult Tadmor (1986a)).
(2) We note that the family of admissible entropy pairs, (2.16), (2.17),
(2.18), becomes smaller once we seek further symmetrization of the
viscous Navier–Stokes terms along the lines of (2.11) (consult Hughes
et al. (1986)).
(3) Finally, we call attention to the fact that, with the particular choice of
entropy pair (U, F ) = (−ρS, −mS) (consult (2.24), (2.25)), the corre-
sponding potential pair (φ, ψ) turns out to be the density and momen-
tum components of the flow, (φ(v), ψ(v)) = (γ − 1)(ρ, m). Hence, in
view of (2.8), (2.9), Euler equations can be rewritten in the intriguing
form
∂ ∂
[∇v ρ] + [∇v m] = 0. (2.26)
∂t ∂x
2
The superscript + (respectively −
) denotes the positive (respectively negative) part of
the indicated scalar.
Entropy stability 461
We close the section with the promised discussion on the entropy stability
of the scalar case. We start with the following result, extending the pen-
etrating scalar arguments of Kružkov (1970), which demonstrates how the
‘richness’ of the family of entropy pairs is converted into a stability state-
ment.
Theorem 2.2. (Tadmor 1997, Theorem 2.1) Assume the system (2.1)
is endowed with an N -parameter family of entropy pairs, (U (u; c), F (u; c)),
c ∈ RN , satisfying the symmetry property
U (u; c) = U (c; u), F (u; c) = F (c; u). (2.27)
Let u1 , u2 be two entropy solutions of (2.1). Then the following a priori
estimate holds:
Z Z
U (u (x, t); u (x, t)) dx ≤ U (u1 (x, 0); u2(x, 0)) dx.
1 2
(2.28)
x x
Let us point out that the elegance of the last result is confronted with the
difficulty of satisfying the symmetry property (2.27). Thus, for example,
N × N symmetric hyperbolic systems are endowed with the N -parameter
family of entropies U (u, c) = |u−c|2/2, but (2.27)
Ru fails for the corresponding
entropy fluxes, F (u, c) = hu − c, f (u)i − hf (w), dwi. The favourable
situation occurs in the scalar case where each convex U serves as an entropy
function. In particular, Kružkov (1970) set the one-parameter family
The symmetry requirement (2.27) holds and (2.28) leads to the following
L1-stability estimate.
Thus there exists a unique (entropy) solution operator associated with the
scalar conservation law (2.1), S(t) : u(·, 0) 7→ u(·, t), which is conservative
and, according to Corollary 2.3, is also L1 -contractive, and hence by the
Crandall–Tartar lemma (Crandall and Tartar 1980), S is order-preserving,
u2(·, 0) ≥ u1 (·, 0) =⇒ S(t)u2(·, 0) ≥ S(t)u1(·, 0). There is a parallel discrete
theory for so-called monotone schemes which respect a similar discrete prop-
erty of order preserving. The entropy stability of such schemes goes back to
the pioneering work of Harten et al. (1976). We will not be able to expand
on the details in the limited framework of this review, but let us mention the
elegant approach of Crandall and Majda (1980), which clarified the entropy
stability of monotone schemes in terms of a comparison with the constant
solution. Sanders (1983) generalized the result to variable grids and we re-
fer to Godlewski and Raviart (1996), Kröner (1997), Tadmor (1998) and
LeVeque (2002) and the references therein for a series of later works, with
particular emphasis on multidimensional extensions. Monotone schemes are
at most first-order accurate (Harten et al. 1976); indeed, being entropy-stable
with respect to all convex entropies, monotone schemes are necessarily lim-
ited to first-order accuracy (Osher and Tadmor 1988). This limitation led
to systematic development of high-resolution schemes which circumvent this
first-order limitation. For a brief overview of the convergence analysis of
such schemes, we refer to Tadmor (1998). Our discussion below focuses on
the question of entropy stability of such first-order as well as higher-order
resolution schemes, in the context of both scalar and systems of conservation
laws.
Entropy stability 463
is solely due to the flux through the local neighbourhoods of the arbitrary
boundaries at x−L and xR .
We are concerned here with the question of entropy stability of such
schemes. To this end, let (U, F ) be an entropy pair associated with the
system (2.1). We ask whether the scheme (3.1) is entropy-stable with re-
spect to such a pair, in the sense of satisfying a discrete entropy inequality
analogous to (2.13), that is,
d 1 h i
U (uν (t)) + Fν+ 1 − Fν− 1 ≤ 0. (3.3)
dt ∆xν 2 2
If, in particular, equality holds in (3.3), we say that the scheme (3.1) is
entropy-conservative.
The answer to this question of entropy stability provided in Tadmor (1987b)
consists of two main ingredients: (i) the use of the entropy variables and (ii)
the comparison with appropriate entropy-conservative schemes. We con-
clude this section with a brief overview.
By making the changes of variables uν = u(vν ), the scheme (3.1) recasts
into the equivalent form
d 1 h i
uν (t) = − gν+ 1 − gν− 1 , uν (t) = u(vν (t)), (3.5)
dt ∆xν 2 2
464 E. Tadmor
1 hD E i 1 hD E i
= ∆vν+ 1 , gν+ 1 − ∆ψν+ 1 + ∆vν− 1 , gν− 1 − ∆ψν− 1
2 2 2 2 2 2 2 2
(Tadmor 1987b, Section 4; see also Osher (1984)). Here ∆ψν+ 1 := ψ(vν+1)−
2
ψ(vν ) denotes the difference of entropy flux potential, (2.9), of two neigh-
bouring grid values vν and vν+1 . Thanks to (2.9), Fν+ 1 is a consistent
2
entropy flux and this brings us to the next result.
and, respectively,
D E
∆vν+ 1 , gν+ 1 = ∆ψν+ 1 . (3.11)
2 2 2
which reveals the role of Qν+ 1 as the numerical viscosity coefficient (e.g.,
2
Tadmor (1984b)).
According to (3.11), scalar entropy-conservative schemes are uniquely de-
∗
termined by the numerical flux gν+ 1 = gν+ 1 , that is,
2 2
∆ψν+ 1 Z 1
2
∗ 2
gν+ 1 := ≡ g vν+ 1 (ξ) dξ,
2 ∆vν+ 1 ξ=− 12 2
2
1
vν+ 1 (ξ) := (vν + vν+1 ) + ξ∆vν+ 1 . (4.4)
2 2 2
Noting that
Z 1
∗
2 d
gν+ 1 = (ξ) · g vν+ 1 (ξ) dξ, (4.5)
2 ξ=− 12 dξ 2
Z 1
2
Q∗ν+ 1 = 2ξg 0 vν+ 1 (ξ) dξ. (4.6)
2 ξ=− 12 2
1
=− f (uν+1 ) − f (uν−1 )
2∆xν
1 h ∗ i
+ Qν+ 1 ∆vν+ 1 − Q∗ν− 1 ∆vν− 1 . (4.7)
2∆xν 2 2 2 2
The entropy stability portion of Theorem 3.1 can now be restated in the
following form.
Corollary 4.1. (Tadmor 1987b, Theorem 5.1) The conservative scheme
(4.7) and (4.3) is entropy-stable, if – and for three-point schemes (p = 1)
only if – it contains more viscosity than the entropy-conservative one (4.6),
3
We use primes to indicate differentiation with respect to primary dependent variables,
e.g., g0 = gv (v), f 00 = fuu (u), etc.
466 E. Tadmor
that is,
Q∗ν+ 1 ≤ Qν+ 1 . (4.8)
2 2
where the supremum is taken over all increasing v = v(u). This yields
Godunov’s viscosity coefficient (Osher 1985, Tadmor 1984b)
" #
f (u ν ) + f (u ν+1 ) − 2f (u)
QGν+ 12
= max . (4.10)
min(uν ,uν+1 )≤u≤max(uν ,uν+1 ) ∆uν+ 1
2
Thus, the scalar schemes which are uniformly entropy-stable with respect
to all convex entropies are precisely those that contain at least as much
numerical viscosity as the Godunov scheme does. These so-called E-schemes
were first identified in Osher (1984); see also Tadmor (1984b).
468 E. Tadmor
The E-schemes are only first-order accurate: consult, e.g., Lemma 4.5
below. Corollary 4.1 enables us to verify the entropy stability of second-
order accurate schemes as well. To this end we recall from (4.6) that
Z 1
2 1
Q∗ν+ 1 = 2ξg 0 vν+ 1 (ξ) dξ, vν+ 1 (ξ) = (vν + vν+1 ) + ξ∆vν+ 1 .
2 ξ=− 1 2 2 2 2
2
Z
d 0
1
2 1
= − ξ2 g vν+ 1 (ξ) dξ, (4.11)
ξ=− 12 4 dξ 2
and hence the entropy-conservative viscosity coefficient Q∗ν+ 1 takes the form
2
Z 1
2 1
Q∗ν+ 1 = − ξ g 00 vν+ 1 (ξ) dξ · ∆vν+ 1 .
2
(4.12)
2 ξ=− 12 4 2 2
Lemma 4.5. Consider the conservative schemes (4.3) with viscosity coef-
ficient, Qν+ 1 , such that Qν+ 1 /∆vν+ 1 is Lipschitz-continuous. Then these
2 2 2
schemes are second-order accurate, in the sense that their local truncation
error is of the order
h i
O |xν+1 − xν |2 + |xν − xν−1 |2 + |xν+1 − 2xν + xν−1 | .
other high-resolution schemes (Majda and Osher 1978, Majda and Osher
1979, Harten 1983a, Harten and Hyman 1983, Nessyahu and Tadmor 1990).
We remark that the careful calculations required in those derivations are
due to the delicate balance of the cubic order of entropy loss, which should
match the third-order dissipation in this case.
In the rarefaction case, ∆uν+ 1 > 0, the integral on the right approximated
2
from below by the midpoint rule; in the case of a shock, ∆uν+ 1 < 0, signs
2
are reversed and we can instead use the trapezoidal rule. Thus we derive a
second-order accurate entropy stable scheme (4.1), whose simple numerical
470 E. Tadmor
Difference schemes that admit the viscosity form (5.1) are precisely the
so-called essentially three-point schemes (Harten (1983a), Tadmor (1987b,
Lemma 5.1)), namely, difference schemes whose numerical flux, f (·, ·, · · ··)
satisfies the restricted consistency relation
f (uν−p+1 , . . . , uν = uν+1 = u, . . . , uν+p ) = f (u).
This is the case of (5.1), with
1 1
f (. . . , uν , uν+1 , . . .) = f (uν ) + f (uν+1 ) − Qν+ 1 (vν+1 − vν ),
2 2 2
vν = v(uν ).
A couple of remarks are in order.
1
vν+ 1 (ξ) := (vν + vν+1 ) + ξ∆vν+ 1 . (5.2)
2 2 2
472 E. Tadmor
1 Z 1 dvν+ 1 (ξ)
2 2
2
= ξg vν+ 1 (ξ) − ξg v v ν+ 1 (ξ) dξ
2 ξ=− 12 ξ=− 12 2 dξ
Z 1
1 2
= f (uν ) + f (uν+1 ) − ξB vν+ 1 (ξ) dξ∆vν+ 1 . (5.3)
2 ξ=− 1 2 2
2
Z 1
2
Q∗ν+ 1 := 2ξB vν+ 1 (ξ) dξ, B(v) = gv (v). (5.5)
2 ξ=− 12 2
The entropy stability portion of Theorem 3.1 can now be conveniently inter-
preted as follows.
Indeed, we can provide a precise measure for the amount of entropy dissi-
pation in terms of the dissipation matrix Dν+ 1 ≡ Dν+ 1 (v(t)) := Qν+ 1−Q∗ν+ 1
2 2 2 2
Entropy stability 473
1 hD E 1D Ei
=− ∆vν− 1 , Dν− 1 ∆vν− 1 + ∆vν+ 1 , Dν+ 1 ∆vν+ 1
4∆xν 2 2 2 4 2 2 2
1D ∗
E 1
Fν+ 1 = vν + vν+1 , gν+ 1 − ψ(vν ) + ψ(vν+1)
2 2 2 2
1 D E
− vν + vν+1 , Dν+ 1 ∆vν+ 1 . (5.8)
4∆xν 2 2
d 1
uν (t) = − f (uν+1 ) − f (uν−1 )
dt 2∆xν
1 h i
+ pν+ 1 ∆uν+ 1 − pν− 1 ∆uν− 1 . (5.9)
2∆xν 2 2 2 2
(5.10)
and hence the viscous part of the scheme (5.9) can be interpreted in terms
of the entropy variables (rather than the conservative ones), as
pν+ 1 ∆uν+ 1 = Qν+ 1 ∆vν+ 1 , (5.11)
2 2 2 2
where
Z 1
2
Qν+ 1 = pν+ 1 H(ξ) dξ, H(ξ) ≡ H vν+ 1 (ξ) . (5.12)
2 2
ξ=− 12 2
474 E. Tadmor
Hence (5.13) holds and entropy stability follows, for any scalar pν+ 1 satis-
2
fying
h i
1 1
pν+ 1 ≥ max 2ξλ H − 2 (ξ)A(ξ)H 2 (ξ)
2 λ,|ξ|≤ 12
h i
= max λ A u vν+ 1 (ξ) . (5.15)
1 2
λ,|ξ|≤ 2
4
Here and below, λk [·] denotes the kth eigenvalue of a matrix.
Entropy stability 475
1 1 1
Since H − 2 (ξ)rk (ξ) are the eigenvectors of the matrix H − 2 (ξ)A(ξ)H 2 (ξ),
and since, by (5.14),
1 1 1 1
H − 2 (ξ)A(ξ)H 2 (ξ) ≡ H − 2 (ξ)B(ξ)H − 2 (ξ), B(ξ) ≡ B vν+ 1 (ξ) , (5.17)
2
1
is a symmetric matrix, it follows after normalization that {H − 2 (ξ)rk (ξ)}
form an orthonormal system, that is,
D 1 1
E
H − 2 (ξ)rk (ξ), H − 2 (ξ)rj (ξ) = δjk . (5.18)
1
We expand H 2 (ξ)∆vν+ 1 and substitute the expansion into the right-hand
2
side of (5.16) to find
D E N Z
D E
1
X 2 2
∆vν+ 1 , Q∗ν+ 1 ∆vν+ 1 = 2ξak (ξ)· rk (ξ), ∆vν+ 1 dξ. (5.19)
2 2 2
2
k=1 ξ=− 12
We compute
d D E
2
D E D E
2
ak (ξ) · rk (ξ), ∆vν+ 1 = ∇v ak (ξ), ∆vν+ 1 · rk (ξ), ∆vν+ 1
dξ 2 2 2
and
d D E
2
D E D E
rk (ξ), ∆vν+ 1 = 2 · rk (ξ), ∆vν+ 1 · ∆vν+ 1 , ∇v rk (ξ)∆vν+ 1 ,
dξ 2 2 2 2
3
and since both terms are of order O ∆vν+ 1 , it follows that the quantity
2
on the right of (5.20) does not exceed
D E
∆v 1 , Q∗ 1 ∆v 1 ≤ C 1 · ∆v 1 3. (5.21)
ν+ ν+
2
ν+ 2
ν+ ν+ 2 2
2
1
0< · IN ≤ Hν+ 1 ≤ K · IN .
K 2
Proof. Using (5.10), the scheme (5.24) meets the desired form in (5.22)
with
Z 1
2
Qν+ 1 = Pν+ 1 H, H ≡ Hν+ 1 = H vν+ 1 (ξ) dξ.
2 2 2
ξ=− 12 2
According to (5.27), the eigenvalues of the matrix on the left are bounded
from below by γν+ 1 · |∆uν+ 1 |. Hence, by (5.25), (5.26), the same is true for
2 2
the eigenvalues of the first matrix on the right; more precisely, we have
h 1
1
i
λ H − 2 Re Qν+ 1 H − 2 ≥ γν+ 1 − Kδν+ 1 · ∆uν+ 1 .
2 2 2 2
1
Multiplying on both sides by H we find, on account of (5.28),
2
h i 1
λ Re Qν+ 1 ≥ γν+ 1 − δν+ 1 · ∆uν+ 1 ≥ Cν+ 1 · ∆vν+ 1 ,
2 K 2 2 2 2 2
matrix for the Euler equations (2.15); Harten and Lax (1981) have shown
its existence in the general case, namely
−1
Aν+ 1 = Bν+ 1 · Hν+ 1, (5.31)
2 2 2
Given a Roe matrix, the viscosity coefficient in (5.24), Pν+ 1 , is then set
2
to be
Pν+ 1 = p Aν+ 1 . (5.33)
2 2
stable schemes. We start by discussing the pros and cons of some first-order
choices.
The original choice of Roe (1981) employs the viscosity function
p(ak ) = |ak |. (5.36)
It has the desirable property that discrete steady shocks are perfectly resolved
on the grid. With this choice, the entropy stability requirement (5.35) reads
|ak | ≥ KCν+ 1 · ∆uν+ 1 ,
2 2
and it is fulfilled as long as we are away from sonic points. Yet this require-
ment may be violated in sonic neighbourhoods where ak ≈ 0; indeed, Roe’s
scheme is the canonical example of an entropy-unstable scheme, for it admits
steady expansion shocks. Theorem 5.7 suggests a simple modification – first
proposed by Osher (1985, Theorem 3.3) – in order to avoid such instability.
Example 5.8. (First-order entropy fix of the Roe scheme) The Roe
scheme (5.24), (5.33), (5.34), is entropy-stable with a viscosity function
n o
p(ak ) = max |ak |, KCν+ 1 · ∆uν+ 1 . (5.37)
2 2
The slightly more viscous modification of Harten (1983a) takes the form
p(ak ) = max |ak |, ε , ∆uν+ 1 1. (5.38)
2
By virtue of (5.39) we can obtain rather detailed information about the en-
tropy dissipation rate of the Roe-type schemes (5.24), (5.33), (5.34).
Let ∆ak (uν ) denote the jump in the kth eigenvalue
∆ak (uν ) = λk (A(uν+1 )) − λk (A(uν )).
In the Appendix we prove the following theorem.
Theorem 5.9. The conservative Roe-type scheme (5.24), (5.33), (5.34) is
480 E. Tadmor
This choice of viscosity function was suggested in Harten and Hyman (1983,
Appendix A) and numerical simulations were carried out in Kaddouri (1993),
for example. To gain a better insight into this choice, we shall distinguish
between three different cases.
2
Case I. ∆ak (uν ) ≤ −Const∆u 1 + O ∆u 1 .
ν+ 2 ν+ 2
Entropy stability 481
In this case the jump ∆ak is dominated by a k-shock, and with sufficiently
small variation, no additional viscosity is required in (5.36), i.e., (5.41) is
reduced to p(ak ) = |ak |. Thus, (5.41) retains the perfect resolution of (suf-
ficiently weak) discrete steady shocks.
2
Case II. ∆ak (uν ) = O ∆uν+ 1 .
2
In this case the jump ∆ak is essentially due to the k-contact field and/or
the balance between the other fields. Here, a minimal
amount
2 of viscosity
is required near sonic points p(ak ) = |ak | + Const ∆uν+ 1 .
2
Case III. Finally, in all other cases we shall identify the jump ∆ak (uν )
as dominated by a k-rarefaction, and as expected, O ∆uν+ 1 amount of
2
dissipation is required near sonic points, p(ak ) = |ak | + Const∆u 1 . ν+ 2
c 2
. (5.44)
2 2 2 ∆vν+ 1 , uν+ 1
2 2
We note that the quantity on the right is well defined since the denominator
does not vanish h(H −1)∗ u, ui > 0. The correction preserves second-order
accuracy, and Lerat and his co-workers (Khalfallah and Lerat 1989) report
on successful applications of such entropy correction in numerical simula-
tions of fluid dynamics problems. Let us point out two limitations to the
present approach: (i) we need to compute the entropy-conservative term
P ∗ u = Q∗ v, which might not be readily available, and (ii), as before, the
entropy correction does not distinguish between different waves within the
same cell. Both points are addressed in the context of the new entropy-
conservative schemes introduced in the next section: consult (6.12) below,
for example.
j k
2
j N
`ν+ 1 , rν+ 1 = δjk . Next, we introduce the intermediate states, vν+ 1 j=1 ,
2 2 2
1
starting with vν+ 1 = vν , and followed by
2
D E
j+1 j j
vν+ 1 = vν+ 1 + `ν+ 1
, ∆v ν+ 1 rjν+ 1 , j = 1, 2, . . ., N, (6.1)
2 2 2 2 2
∗
with a numerical flux gν+ 1 given by
2
j+1 j
XN ψ v 1 − ψ v 1
∗ ν+ ν+ 2 j
gν+ 1 =
j 2 `ν+ 1 , (6.3)
2
j=1 `ν+ 1 , ∆vν+ 1 2
2 2
N
X
j+1 j
= ψ vν+ 1 − ψ v ν+ 1 2 2
j=1
N +1 1
= ψ vν+ 1 − ψ vν+ 1 = ∆ψν+ 1 .
2 2 2
484 E. Tadmor
Z 1 j+ 1 D E
2
j
= 2
g vν+ 1 (ξ) dξ, rν+ 1 `jν+ 1 , ∆vν+ 1 .
ξ=− 12 2 2 2 2
(6.5)
Inserting this into (6.3), we find that the entropy-conservative flux can be
equivalently written as
XN Z 1 j+ 1
2
∗
gν+ 1 = g vν+ 1 (ξ) dξ, rν+ 1 `jν+ 1 ,
2 j
(6.6)
2
j=1 ξ=− 12 2 2 2
n j+ 12
oN +1
Remark. We note that if we let rν+ 1 collapse into the same direc-
2 j=1
tion of ∆vν+ 1 , then the new entropy-conservative flux (6.5) collapses into
2
the entropy-conservative flux of the ‘first kind’ studied earlier in Section 5.
As before, the new entropy-conservative schemes admit a viscosity form,
subject to the phase space path. Considering a typical subpath factor on the
right of (6.6), we integrate by parts along the lines of (5.3), to obtain
Z 1 D j+ 1 E
2 d 2 j
(ξ) g vν+ 1 (ξ) dξ, rν+ 1
ξ=− 12 dξ 2 2
1D j
j+1
j
E
= f uν+ 1 + f uν+ 1 , rν+ 1
2 2 2 2
Z 1 D ED E
2 j+ 1
+ ξ rjν+ 1 , B vν+ 21 (ξ) rjν+ 1 `jν+ 1 , ∆vν+ 1 .
ξ=− 12 2 2 2 2 2
j+ 1
(i) Comparison. We seek appropriate viscosity amplitudes, qν+ 21 , which
2
upper-bound the amount of entropy-conservative viscosities on each subpath
j+ 1
in phase space, vν+ 21 (ξ), so that (compare Corollary 5.1)
2
D j+ 1 ,∗
E j+ 1
rjν+ 1 , Qν+21 rjν+ 1 ≤ qν+ 21 . (6.9)
2 2 2 2
A straightforward argument along the lines of our previous results yields the
following result.
486 E. Tadmor
1 X j+ 12 D j E
N
+ qν+ 1 `ν+ 1 , ∆vν+ 1 `jν+ 1
2∆xν 2 2 2 2
j=1
N
X j+ 1
D E
− qν− 21 `jν− 1 , ∆vν− 1 `jν− 1 , (6.10)
2 2 2 2
j=1
(ii) Choice of path. The new ingredient here is the choice of a proper sub-
path in phase space. We demonstrate the advantage of using such a subpath
in the context of second-order accurate reformulation of the conservative
schemes outlined in Corollary 6.2. Let
n j+ 1 o
wk (v(ξ)) = wk vν+ 21 (ξ)
2
j+ 1
be the orthonormal eigensystem of the symmetric B = B vν+ 21 (ξ) ,
2
j+ 1
j+ 1
B vν+ 21 (ξ) wk (v(ξ)) = bk (v(ξ))wk(v(ξ)), bk (v(ξ)) := λk B vν+ 21 .
2 2
j P
j
Expanding rν+ 1 = k wk (v(ξ)), rν+ 1 wk (v(ξ)), we rewrite the amount of
2 2
entropy-conservative viscosity corresponding to a typical subpath on the left
of (6.9)
D j+ 1 ,∗
E Z 1
2
D j+ 1 E
rjν+ 1 , Qν+ 21 rjν+ 1 = 2ξ rjν+ 1 , B vν+ 21 (ξ) rjν+ 1 dξ
2 2 2 ξ=− 12 2 2 2
N Z
X 1 D E2
2
= 2ξbk (v(ξ)) wk (v(ξ)), rjν+ 1 dξ.
k=1 ξ=− 12 2
Simple upper bounds, for instance, 2ξbk (v(ξ)) ≤ supξ |bk (v(ξ))|, characterize
the first-order Roe-type schemes. For second-order accuracy, we perform one
Entropy stability 487
Here,
second-order
accuracy is reflected by viscosity amplitudes of order
O ∆vν+ 1 along each subpath (being entropy-conservative, the amount of
2
entropy dissipation is zero). How should we choose an appropriate subpath?
To simplify matters we consider the symmetric case where the entropy and
N
conservative variables coincide, B(v) = A(u). We let ujν+ 1 j=1 be the
2
breakpoints along the path of (approximate) solutions to the Riemann prob-
lem. It is well known (Lax 1957) that each subpath is directed along the
eigensystem of A ujν+ 1 , that is, uj+1
ν+ 1
− ujν+ 1 ∼ rjν+ 1 , so wk ∼ rkν+ 1 is the
2 2 2 2 2
normalized eigensystem of A. With this choice, all but one of the terms on
the right of (6.11) vanish to higher order (in |∆uν+ 1 |) and the leading term
2
governing entropy dissipation is given by
D j+ 1 ,∗
E Z 12 1 D j+ 1 E
rjν+ 1 , Qν+21 rjν+ 1 ≈ − ξ 2 ∇uaj uν+ 21 (ξ) , rjν+ 1 dξ.
2 2 2 ξ=− 1 4 2 2
2
1 X h j+1 i+
N
+ aj uν+ 1 − aj ujν+ 1 `jν+ 1
8∆xν 2 2 2
j=1
N h
X i+
− aj uj+1
ν− 1 − a j u j
ν− 1 `jν− 1 . (6.14)
2 2 2
j=1
We note in passing that use of the dissipation matrix Dν+ 1 is restricted here
2
to entropy-conservative schemes of the ‘first kind’ discussed in Section 5,
and we can use a similar, refined argument with the entropy-conservative
schemes of the ‘second kind’ in Section 6, leading to the corresponding gen-
eralization of the fully discrete entropy stability analysis presented below.
1
To discretize in time, we introduce a local time step, tn+1 = tn + ∆tn+ 2 .
We shall use superscripts to denote dependence on the time level, for in-
stance, unν = u(v(xν , tn )), gν+
n n
1 = gν+ 1 (v(t )), etc. To simplify notation,
2 2
we suppress the variability of the time step and grid cell width, abbreviating
1
∆tn+ 2 /∆xν = ∆x
∆t
. We shall study the entropy stability of the fully discrete
schemes in terms of three prototype examples, which demonstrate the balance
between the entropy dissipation from spatial stencil vs. the entropy dissipa-
tion/production due to the time discretization. We begin with the following.
∆t (x) n+1 1
− E v − EνBE vn+ 2
∆x ν
∆t (x) n+1 1
=− Eν v − EνBE vn+ 2 ≤ 0. (7.11)
∆x
Entropy stability is enhanced by fully implicit time discretization. In con-
trast, explicit time discretization, discussed in the next example, leads to
entropy production. Thus, the entropy stability of explicit schemes hinges
on a delicate balance between temporal entropy production and spatial en-
tropy dissipation.
Example 7.2. (Explicit forward Euler (FE) time discretization)
We discretize (7.1) by the forward Euler scheme
∆t h i
un+1
ν = u n
ν − g 1 (v n
) − g 1 (v n
) . (7.12)
∆x ν+ 2 ν− 2
Entropy stability 491
Now, the identity (7.7), (7.8) can be put into the equivalent form
n n+1 1
vν , uν − unν = U un+1 ν − U unν − EνF E vn+ 2 , (7.13)
1
with entropy production EνF E (vn+ 2 ) given by
Z 1 D
FE n+ 12
2 1 n+ 1 n+ 1 n+ 1 E
Eν v := + ξ ∆vν 2 , H vν 2 (ξ) ∆vν 2 dξ ≥ 0.
ξ=− 12 2
(7.14)
We multiply (7.12) by vνn , and together with the spatial dissipation of en-
tropy quantified in (7.5), we arrive at
∆t h n i ∆t (x) n 1
U un+1
ν − U (u n
ν ) + F ν+ 1 − F n
ν− 1 =− Eν (v ) + EνF E vn+ 2 .
∆x 2 2 ∆x
(7.15)
To study the entropy stability of (7.12), we therefore need to upper-bound
the entropy production EνF E , in terms of the spatial dissipation matrices
Dν± 1 , which are responsible for the entropy dissipation in (7.5). We proceed
2
as follows. From (7.14) we have
Z 1 D
1 2 1 n+ 1 n+ 1 n+ 1 E
EνF E vn+ 2 ≤ + ξ ∆vν 2 , H vν 2 (ξ) ∆vν 2 dξ
ξ=− 1 2
2
K n+ 1 2 K 3 n+ 12 2
≤ ∆vν 2 ≤ ∆uν , (7.16)
2 2
where K 2 is the condition number of H: see (5.26). To upper-bound the time
n+ 12 n ) − g(v n ) = B n
differences, ∆uν , we recall g(vν+1 ν ν+ 12 ∆vν+ 1 with Bν+ 12
R n
2
given in (5.32) as Bν+ 1 = B vν+ 1 (ξ) dξ. This enables us to rewrite the
2 2
discrete forward Euler scheme (7.12) in the equivalent incremental form
∆t h
un+1
ν − u n
ν = g v n
ν+1 − g v n
ν
n
+ Qν+ 1 ∆vν+ 1
2∆x 2 2
i
+ g vνn − g vν−1 n n
+ Qν− 1 ∆vν− 1
2 2
∆t h
n
n
i
= Bν+ 1 + Qν+ 1 ∆vν+ 1 + Bν− 1 + Qν− 1 ∆vν− 1 .
2∆x 2 2 2 2 2 2
Finally, we recall the viscosity matrix Qν+ 1 = Q∗ν+ 1 + Dν+ 1 . This enables
2 2 2
us to rewrite the last expression as
∆t h e
e 1 + D 1 ∆vn 1 ,
i
uνn+1 − unν = n
Bν+ 1 + Dν+ 1 ∆vν+ 1 + B ν− 2 ν− 2 ν− 2
2∆x 2 2 2
(7.17)
492 E. Tadmor
where
Z
1
2
e := B + Q∗ =
B n
(1 + 2ξ)B vν+ 1 dξ = Bν+ 1 + O ∆vν+ 1 .
ξ=− 12 2 2 2
Compared with the spatial entropy dissipation in (7.5), we find that the
1
∆t (x) n
forward Euler scheme is entropy-stable, − ∆x Eν (v ) + EνF E (vn+ 2 ) ≤ 0,
provided D is sufficiently large that
2 2
3 ∆t e 1 + D 1 ≤ ∆t D 1 .
K B ν+ 2 ν+ 2 (7.19)
∆x ∆x ν+ 2
We consider the two prototype examples of centred and upwind schemes.
∆x
If we set Dν+ 1 = 2∆t IN ×N we obtain the centred modified Lax–Friedrichs
2
(MLxF) scheme (e.g., Tadmor (1984b))
1 n ∆t
un+1
ν = uν+1 + 2unν + unν−1 + f unν+1 − f unν−1 . (7.20)
4 2∆x
To simplify matters, we consider the symmetric case, where the Bs are
turned into As, and (7.19) with condition number K = 1 yields the entropy
stability of the MLxF for sufficiently small CFL number
√
∆t e 2−1
max |λ(A)| ≤ .
∆x λ 2
e 1 | leads to the
Similarly, the viscosity coefficient matrix, Dν+ 1 = |A
2
ν+ 2
upwind scheme
∆t
un+1
ν = unν − f unν+1 − f unν−1
2∆x
1 h
+ Aν+ 1 + Q∗ν+ 1 + Q∗ν+ 1 ∆vν+n
1
2∆xν 2 2 2 2
i
∗
− A 1 + Q 1 + Q 1 ∆v 1 .∗ n
(7.21)
ν− 2 ν+ 2 ν− 2 ν− 2
(1) CFL optimality. In both examples of the centred and upwind schemes,
entropy stability is obtained under less than optimal CFL conditions,
which is due to less than optimal bounds on the entropy production
rate, EνF E . In particular, the resulting entropy stability condition (7.19)
excludes the entropy stability of second-order fully discrete schemes,
which are identified with Lax–Wendroff (LxW) dissipation matrices of
∆t 2
order Dν+ 1 ∼ ∆x Ãν+ 1 .
2 2
1
Noting that vn+ 2 , un+1 − un = U (un+1 ) − U (un ), we conclude the follow-
ing.
∆t ∆t
un+1
ν+ 1
:= unν − f unν+1 − f unν + Pν+ 1 unν+1 − unν ,
2 ∆x ∆x 2
∆t ∆t
un+1
ν− 1
:= unν − f unν − f unν−1 − Pν+ 1 unν − unν−1 ,
2 ∆x ∆x 2
and we study the entropy inequality for each term. This decomposition
Entropy stability 495
into left- and right-handed stencils in the context of cell entropy inequal-
ity was first introduced in Tadmor (1984b). We begin by considering un+1
ν+ 1
.
2
To this end, we set unν+ 1 (s) := unν +s(unν+1 −unν ) and the following inequality
2
is sought (here and below, ∆u := unν+1 − unν ):
∆t
I+ := U un+1
ν+ 12
− U unν + F unν+1 − F unν
∆x
Z 1 D E
∆t
− U 0 unν+ 1 (s) , Pν+ 1 ∆u ds ≤ 0. (8.2)
∆x s=0 2 2
Z 1 D E
∆t
= U 0 unν+ 1 (s) A unν+ 1 (s) , ∆u ds.
∆x s=0 2 2
∆t E
− Pν+ 1 − A unν+ 1 (s) ∆u ds. (8.3)
∆x 2 2
Next, we introduce
unν+ 1 (r, s) := unν+ 1 (s)+r unν −unν+ 1 (s) ≡ unν +s(1−r) unν+1 −unν , (8.4)
2 2 2
496 E. Tadmor
and we set
∆t n
un+1
ν+ 1
(r, s) = unν+ 1 (r, s) − f uν+ 1 (s) − f unν+ 1 (r, s)
2 2 ∆x 2 2
∆t
+ P 1 un 1 (s) − unν+ 1 (r, s)
∆x ν+ 2 ν+ 2 2
so that un+1
ν+ 1
(0, s) = unν+ 1 (s) and un+1
ν+ 1
(1, s) = un+1
ν+ 1
(s). This then yields
2 2 2 2
Z 1
d 0 n+1
U 0 un+11 (s) − U
ν+ 2
0
u n
ν+1 (s) = U u 1 (r, s) dr
ν+ 2
r=0 dr
Z 1 ∆t ∆t
= −s U 00 un+1
ν+ 12
(r, s) dr I + A u n
ν+ 12
(r, s) − P 1 ∆u.
r=0 ∆x ∆x ν+ 2
Inserting the last expression into the right-hand side of (8.3) we end up with
Z 1
∆t n ∆t
I+ = − s I+ A uν+ 1 (r, s) − P 1 ∆u,
r,s=0 ∆x 2 ∆x ν+ 2
∆t ∆t
00 n+1 n
U uν+ 1 (r, s) − A uν+ 1 (s) + P 1 ∆u dr ds. (8.5)
2 ∆x 2 ∆x ν+ 2
To continue, we focus our attention on two prototype cases.
(i) The scalar case. The positivity of the last expression on the right of
(8.5) follows from a CFL condition
∆t n ∆t ∆t n
a uν+ 1 (s) ≤ pν+ 1 ≤ I + a uν+ 1 (r, s) . (8.6)
∆x 2 ∆x 2 ∆x 2
Again, the last expression is nonconservative, but together with (8.2) we end
up with the cell entropy inequality.
Corollary 8.1. Consider the fully discrete scalar scheme (8.1) and assume
the CFL condition
∆t n ∆t ∆t n
a uν+ 1 (s) ≤ pν+ 1 ≤ I − a uν+ 1 (r, s) (8.8)
∆x 2 ∆x 2 ∆x 2
Entropy stability 497
Z 1 D #
E
0 n n n
− U uν− 1 (s) , pν− 1 uν − uν−1 ds .
2 2
s=0
n+1
yields the quasi-cell entropy inequality for uν− 1 , and the following conclu-
2
sion.
498 E. Tadmor
Corollary 8.2. Consider the fully discrete scheme (8.1) consistent with
the symmetric system (2.1) and assume the CFL condition5
∆t n ∆t ∆t n
A uν+ 1 (s) ≤ Pν+ 1 ≤ I − A uν+ 1 (r, s) (8.10)
∆x 2 ∆x 2 ∆x 2
5
We recall (consult (5.34)) that |A| stands for the absolute value of A, defined by its
spectral decomposition
|a1 |
.. −1
|A| = R . R .
|aN |
Entropy stability 499
We find that the upwind scheme is entropy-stable for the quadratic en-
tropy function (for symmetric systems) and for all convex entropies (for
scalar equations), provided the CFL condition (8.8), (8.10) holds, which
amounts to
∆t 1
sup λ A uν+ 1 (s) ≤ . (8.12)
∆x s,λ 2 2
Again, a linearized von Neumann stability analysis reveals that the CFL
condition is sharp.
(1) Comparison with Roe scheme. Consider the numerical viscosity of the
Roe-type scheme (5.24), (5.33), Pν+ 1 = |Aν+ 1 |. A comparison with
2 2
the upwind scheme (8.11), P 1 = sups A un 1 (s) , reveals that the
ν+ 2 ν− 2
entropy stability of the latter is explained by taking into account all in-
termediate values between two neighbouring values, uν and uν+1 . An
alternative entropy correction discussed in Example 5.10 adds the ad-
ditional term of order O(|∆u|) to compensate for the missing interme-
diate values.
(2) Extensions. The entropy stability results in this section are based on
a homotopy approach. Our initial point was the essentially three-point
scheme (8.1). The same homotopy approach refinement, starting with
500 E. Tadmor
9. Higher-order extensions
We generalize the construction of second-order entropy-conservative schemes
to higher orders. To this end, we revisit the original derivation of the second-
order entropy-conservative schemes (Tadmor 1986b), using finite element
discretization. We begin with the weak formulation of the systems of con-
servation laws (2.6),
Z Z
∂ ∂
w(x, t), u(v) = w(x, t), g(v) dx dt, Ω ⊂ R × (0, T ),
Ω ∂t Ω ∂x
(9.1)
where w(·) is an arbitrary C0∞ (Ω) test function. The key point is the use
of the entropy variables, which enables us to use the standard finite element
framework where both the primary computed solution v and the test function
w belong to the same finite-dimensional scale of spaces. In particular, let the
Entropy stability 501
P
trial solution v̂ = µ vµ (t)Ĥµ (x) be chosen from the typical finite element
space spanned by the C 0 ‘hat functions’
x−x
µ−1
xµ −xµ−1 , xµ−1 ≤ x ≤ xµ ,
Ĥµ (x) = x −x
µ+1
xµ+1 −xµ , xµ ≤ x ≤ xµ+1 .
Testing (9.1) against w(x) = w(x) = Ĥν (x), the right-hand side of (9.1)
yields
Z xν+1 !
∂ X
Ĥν (x)g vµ (t)Ĥµ(x) dx dt =
xν−1 ∂x µ
"Z 1 Z 1 #
2
2
− g vν+ 1 (ξ) dξ − g vν− 1 (ξ) dξ , (9.2)
2 2
ξ=− 12 ξ=− 12
Equating (9.2) and (9.3) while neglecting the quadratic error term, we end
up with the entropy-conservative scheme (5.2):
Z 1
d 1 h ∗ ∗
i
∗
2
uν (t) = − gν+ 1 − gν− 1 , gν+ 1 = g vν+ 1 (ξ) dξ.
dt ∆xν 2 2 2 ξ=− 1 2
2
∗∗
with a numerical flux, gν+ 1 = g(vν−1 , vν , vν+1 , vν+2 ), given by
2
Z 1
2
∗∗
gν+ 1 = g vν+ 1 (ξ) dξ
2 ξ=− 12 2
1 h ∗∗ i
− Qν+ 3 (vν+2 − vν+1 ) − Q∗∗ 1 (vν − vν−1 ) .
ν− 2
(9.6)
12 2
Here, Q∗∗
ν+ 1
is a secondary viscosity coefficient depending on
2
Q∗∗
ν+ 1
= Q∗∗(vν−1 , vν , vν+1 ) (9.7)
2
REFERENCES
R. Abramov and A. Majda (2002), ‘Discrete approximations with additional con-
served quantities: Deterministic and statistical behavior’, preprint.
R. Abramov, G. Kovačič and A. Majda (2003), ‘Hamiltonian structure and statis-
tically relevant conserved quantities for the truncated Burgers–Hopf equation’,
Comm. Pure Appl. Math. 56, 1–46.
H. Aiso (1993), ‘Admissibility of difference approximations for scalar conservation
laws’, Hiroshima Math. J. 23, 15–61.
A. Arakawa (1966), ‘Computational design for long-term numerical integration of
the equations of fluid motion: Two-dimensional incompressible flow, Part I’,
J. Comput. Phys. 1, 119–143.
S. Bianchini and A. Bressan (2003), ‘Vanishing viscosity solutions of nonlinear
hyperbolic systems’, Ann. Math. To appear.
F. Bouchut (2002), ‘Entropy satisfying flux vector splitting and kinetic BGK mod-
els’, Numer. Math., DOI http://dx.doi.org/10.1007/s00211-002-0426-9.
F. Bouchut, Ch. Bourdarias and B. Perthame (1996), ‘A MUSCL method satisfying
all the numerical entropy inequalities’, Math. Comp. 65, 1439–1461
A. Bressan (2003), Viscosity solutions of nonlinear hyperbolic systems, in Hyper-
bolic Problems: Theory, Numerics and Applications, Proc. 9th Hyperbolic
Conference (T. Hou and E. Tadmor, eds), Springer. To appear.
Entropy stability 503
J. von Neumann and R. D. Richtmyer (1950), ‘A method for the numerical calcu-
lation of hydrodynamic shocks’, J. Appl. Phys. 21, 232–237.
P. Olsson (1995), ‘Summation by parts, projections, and stability, III’, RIACS
Technical report, 95.06.
S. Osher (1984), ‘Riemann solvers, the entropy condition, and difference approxi-
mations’, SIAM J. Numer. Anal. 21, 217–235.
S. Osher (1985), ‘Convergence of generalized MUSCL schemes’, SIAM J. Numer.
Anal. 22, 947–961.
S. Osher and S. Chakravarthy (1984), ‘High resolution schemes and the entropy
condition’, SIAM J. Numer. Anal. 21, 955–984.
S. Osher and F. Solomon (1982), ‘Upwind difference schemes for hyperbolic con-
servation laws’, Math. Comp. 38, 339–374.
S. Osher and E. Tadmor (1988), ‘On the convergence of difference approximations
to scalar conservation laws’, Math. Comp. 50, 19–51.
R. D. Richtmyer and K.W. Morton (1967), Difference Methods for Initial-Value
Problems, Interscience, 2nd edn.
P. L. Roe (1981), ‘Approximate Riemann solvers, parameter vectors and difference
schemes’, J. Comput. Phys. 43, 357–372.
V. V. Rusanov (1961), ‘Calculation of interaction of non-steady shock-waves with
obstacles’, J. Comput. Math. Phys., USSR 1, 267–279.
R. Sanders (1983), ‘On convergence of monotone finite difference schemes with
variable spatial differencing’, Math. Comp. 40, 91–106.
D. Serre (1991), ‘Richness and the classification of quasilinear hyperbolic systems’,
Multidimensional Hyperbolic Problems and Computations, Minneapolis, MN,
Inst. Math. Appl., Vol 29, Springer, pp. 315–333.
D. Serre (1999), Systems of Conservation Laws, 1: Hyperbolicity, Entropies, Shock
Waves (English translation), Cambridge University Press.
M. Schonbek (1982), ‘Convergence of solutions to nonlinear dispersive equations’,
Comm. PDEs 7, 959–1000.
J. Smoller (1983), Shock Waves and Reaction Diffusion Equations, Springer,
Berlin.
T. Sonar (1992), ‘Entropy production in second-order three-point schemes’, Numer.
Math. 62, 371–390.
A. Szepessy (1989), ‘An existence result for scalar conservation laws using measure
valued solutions’, Comm. PDEs 14, 1329–1350.
E. Tadmor (1984a), ‘Skew-adjoint form for systems of conservation form’, J. Math.
Anal. Appl. 103, 428–442.
E. Tadmor (1984b), ‘Numerical viscosity and the entropy condition for conservative
difference schemes’, Math. Comp. 43, 369–381.
E. Tadmor (1986a), ‘A minimum entropy principle in the gas dynamics equations’,
Appl. Numer. Math. 2, 211–219.
E. Tadmor (1986b), Entropy conservative finite element schemes, in Numerical
Methods for Compressible Flows: Finite Difference Element and Volume Tech-
niques, Proc. Winter Annual Meeting of the Amer. Soc. Mech. Eng. AMD,
Vol. 78 (T. E. Tezduyar and T. J. R. Hughes, eds), pp. 149–158.
E. Tadmor (1987a), ‘Entropy functions for symmetric systems of conservation
laws’, J. Math. Anal. Appl. 121, 355–359.
Entropy stability 507
E. Tadmor (1987b), ‘The numerical viscosity of entropy stable schemes for systems
of conservation laws, I’, Math. Comp. 49, 91–103.
E. Tadmor (1989), ‘Convergence of spectral methods for nonlinear conservation
laws’, SIAM J. Numer. Anal. 26, 30–44.
E. Tadmor (1997), Approximate solution of nonlinear conservation laws and related
equations, in Recent Advances in Partial Differential Equations and Applica-
tions, Proc. 1996 Venice Conference in Honor of Peter D. Lax and Louis
Nirenberg on their 70th Birthday (R. Spigler and S. Venakides, ed.), AMS
Proceedings Symp. Appl. Math. Vol. 54, Providence, RI, pp. 321–368.
E. Tadmor (1998), Approximate solutions of nonlinear conservation laws, in Ad-
vanced Numerical Approximation of Nonlinear Hyperbolic Equations, Lecture
Notes from CIME Course Cetraro, Italy, 1997 (A. Quarteroni, ed.), Vol. 1697
of Lecture Notes in Mathematics, Springer, pp. 1–150.
L. Tartar (1975), Compensated compactness and applications to partial differential
equations, in Nonlinear Analysis and Mechanics, Heriot-Watt Symposium,
Vol. 4 (R. J. Knopps, ed.), Vol. 39 of Research Notes in Mathematics, Pitman
Press, pp. 136–211.
A. I. Vol’pert (1967), ‘The spaces BV and quasilinear equations’, Math. USSR-Sb.
2, 225–267.
H. Yang (1996a), ‘On wavewise entropy inequalities for high-resolution schemes, I:
The semidiscrete case’ Math. Comp. 65, 45–67. Supplement Math. Comp. 65
S1–S13.
H. Yang (1996b), ‘Convergence of Godunov type schemes’ Appl. Math. Letters 9,
63–67.
H. Yang (1998), ‘On wavewise entropy inequalities for high-resolution schemes, II:
Fully discrete MUSCL schemes with exact evolution in small time’, SIAM J.
Numer. Anal. 36, 1–31.
Once more (consult Example 5.2), we use (5.10) to rewrite the viscous part
of (A.1) in terms of the entropy variables, obtaining
d 1
uν (t) = − f (uν+1 ) − f (uν−1 )
dt 2∆xν
1 h i
+ Qν+ 1 ∆vν+ 1 − Qν− 1 ∆vν− 1 , (A.2)
2∆xν 2 2 2 2
508 E. Tadmor
where
Z 1
2
Qν+ 1 = p Aν+ 1 H(ξ) dξ, H(ξ) ≡ H vν+ 1 (ξ) . (A.3)
2
ξ=− 12 2 2
Corollary 5.1 suggests that the entropy dissipation of these schemes should
be measured by the quantity h∆vν+ 1 , Qν+ 1 ∆vν+ 1 i. A lower bound for the
2 2 2
latter is provided in the following lemma.
Lemma A1. Let {rk , ak } be the eigensystem of Aν+ 1 , and assume that
2
2
|rk − rk (ξ = 0)| + |ak − ak (ξ = 0)| ≤ Const ∆vν+ 1 . (A.4)
2
Then we have
D E
∆vν+ 1 , Qν+ 1 ∆vν+ 1
2 2 2
N h Z
2 i D E
1
X 2
≥ p(ak ) − Const∆vν+ 1 rk (ξ), ∆v 2 dξ. (A.5)
2
ν+ 21
k=1 ξ=− 12
1
Proof. Using the orthonormal system {H − 2 (ξ)rk (ξ)} in (5.18), we can ex-
pand the right-hand side of the equality (A.3), which we rewrite as
D E Z 12 D 1 1
E
∆vν+ 1 , Qν+ 1 ∆vν+ 1 = ∆vν+ 1 , p Aν+ 1 H 2 (ξ)·H 2 (ξ)∆vν+ 1 dξ,
2 2 2 2 2 2
ξ=− 12
and find
D E XN Z 1 D E
2
∆vν+ 1 , Qν+ 1 ∆vν+ 1 = ∆vν+ 1 , p Aν+ 1 rk (ξ) αk (ξ) dξ,
2 2 2 2 2
k=1 ξ=− 12
(A.6)
where αk (ξ) abbreviates αk (ξ) := rk (ξ), ∆vν+ 1 .
2
Consider the quantities on the right of (A.6): their dependence on ξ is
reflected through their dependence on vν+ 1 (ξ) = 12 (vν + vν+1 ) + ξ∆vν+ 1 ;
2 2
for such quantities we have
s
d X(ξ)
ds
s
.
dξ s ≡ X v ν+ 12 (ξ) = O ∆v ν+ 12 (A.7)
dξ s
By Taylor’s theorem,
ξ2
rk (ξ) = rk (0) + ξ ṙk (0) + r̈k (θξ), for some θ ∈ [0, 1].
2
Here and below, (˙) denotes ξ-differentiation, dξ
d
( ). In view of (A.7), |r̈k | ≤
2
Const∆v 1 , and together with assumption (A.4) we have
ν+ 2
2
rk (ξ) = rk + ξ ṙk (0) + Jk , |Jk | ≤ Const∆vν+ 1 . (A.8)
2
Entropy stability 509
we substitute the last three terms into three summations on the right of
(A.9), respectively, and end up with
D E
∆vν+ 1 , Qν+ 1 ∆vν+ 1
2 2 2
N
X Z N
= p(ak ) · α2k (ξ) dξ
k=1 ξ=− 12
N Z
X 1 D h i E
2
+ ξ ∆vν+ 1 , p Aν+ 1 − p(ak ) · IN ṙk (0) αk (ξ) dξ
1 2 2
k=1 ξ=− 2
XN Z 1 D h i E
2
+ ∆vν+ 1 , p Aν+ 1 − p(ak ) · IN Jk αk (ξ) dξ
2 2
k=1 ξ=− 12
N
X Z 1
2 D E
= p ak rk (ξ), ∆v 2 dξ + II + III . (A.10)
ν+ 21
k=1 ξ=− 12
3
Since Jk αk (ξ) is of order O ∆vν+ 1 , we have
2
4
|III | ≤ ConstIII · ∆vν+ 1 ; (A.11)
2
The result (A.5) now follows, noting that H(ξ) ≤ K · IN (see (5.26)), and
hence
Z 1
1
∆v 1 2 ≤ 1
2
H 2 (ξ)∆v 1 2 dξ
ν+ 2 ν+
K ξ=− 1 2
2
N Z 1
1 X 2 D E
2
≤ rk (ξ), ∆vν+ 1 dξ; (A.13)
K ξ=− 1 2
k=1 2
X N
1 |ak | 2
≤ ∆ak (uν ) + εk |ak | + 1 +
Const ∆vν+ 1
6 εk 2
k=1
Z 1
2 D E
× rk (ξ), ∆v 1 2 dξ. (A.14)
ν+ 2
ξ=− 12
Moreover, we have
ξ2 2
αk (±ξ) = αk (0) ± ξ ȧk (0) + äk , |äk | ≤ Const∆vν+ 1 ;
2 2
we substitute this into the right-hand side of (A.17): since ȧk (0), α2k (0) and
2 3
αk (0)α̇k (0) are of order O ∆vν+ 1 , O ∆vν+ 1 , and O ∆vν+ 1 re-
2 2 2
spectively, we obtain
ak (ξ)α2k (ξ) − ak (−ξ)α2k (−ξ)
4
= 2ξ ȧk (0)α2k (0) + 4ξak (0)αk (0)α̇k (0) + O ∆vν+ 1 . (A.18)
2
Inserting the latter expression into (A.15), we find after integration that
D E
∆vν+ 1 , Q∗ν+ 1 ∆vν+ 1
2 2 2
N
X 1 4
≤ ȧk (0)α2k (0) + 2ak (0)αk (0)α̇k (0) + Const ∆vν+ 1 .
6 2
k=1
X1N
|ak | 2
≤ ȧk (0) + εk |ak | + 1 + Const∆vν+ 1
6 εk 2
k=1
Z 1
2 D E
× rk (ξ), ∆v 1 2 dξ, (A.21)
ν+ 2
ξ=− 12
2
and the result (A.14) follows, noting that ȧk (0) = ∆ak (uν ) + O ∆vν+ 1 .
2