0% found this document useful (0 votes)
11 views

Simulation on porous

Uploaded by

priyamoh2013
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
11 views

Simulation on porous

Uploaded by

priyamoh2013
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 22

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/326004261

Did Numerical Methods for Hyperbolic Problems Take a Wrong Turning?

Chapter · June 2018


DOI: 10.1007/978-3-319-91548-7_39

CITATIONS READS

5 448

1 author:

Philip Roe
University of Michigan
218 PUBLICATIONS 27,356 CITATIONS

SEE PROFILE

All content following this page was uploaded by Philip Roe on 05 October 2024.

The user has requested enhancement of the downloaded file.


J Sci Comput (2017) 73:1094–1114
DOI 10.1007/s10915-017-0555-z

Is Discontinuous Reconstruction Really a Good Idea?

Philip Roe1

Received: 1 February 2017 / Revised: 31 August 2017 / Accepted: 5 September 2017 /


Published online: 12 October 2017
© The Author(s) 2017. This article is an open access publication

Abstract It has been almost automatically assumed for a quarter century that the numerical
solution of hyperbolic conservation laws is best accomplished by making a reconstruction of
the initial data that is only piecewise continuous. The effect of the discontinuities is taken into
account by means of Riemann solvers. This strategy has enjoyed great practical success but
introduces only one-dimensional physics as a guide to the discretization of multidimensional
problems. This article points out some of the resulting defects and proposes an alternative
viewpoint. The chief novelty of the new “Active Flux” method, apart from the elimination of
discontinuities, is the division into advective and acoustic disturbances, with acoustics being
handled by exploiting classical solutions to the scalar wave equation. Results for a standard
test problem for the compressible Euler equations indicate that an order of magnitude increase
in cost effectiveness may be possible by following this approach, compared to the traditional
one.

Keywords Hyperbolic conservation laws · Numerical advection · Compressible flow ·


Wave equations

1 Some History, By Way of Introduction

In 1959, there appeared in the Russian mathematics journal Matematicheskii Sbornik, (Math-
ematical Miscellany) a paper that has exerted, directly or indirectly, a profound influence
on almost all of the methods subsequently used to find numerical solutions of hyperbolic
conservation laws. These include the so-called Godunov-type finite-volume methods and

In honor of Professor Shu’s 60th birthday.

This research was supported in part by NASA Cooperative Agreement NNX12AJ70A.

B Philip Roe
[email protected]
1 Department of Aerospace Engineering, University of Michigan, Ann Arbor, MI 48109, USA

123
J Sci Comput (2017) 73:1094–1114 1095

their successors, such as discontinuous Galerkin and WENO schemes. In his paper [13]
Sergei Godunov showed that no numerical method for the linear scalar advection equation
∂t u + a∂x u = 0 can be both monotone and better than first-order accurate. He also showed
that the linear monotone advection scheme with smallest truncation error was the simplest
upwind scheme, and offered a plausible generalization to one-dimensional systems which
is now well-known as Godunov’s method. This consists of projecting the initial data into
the space of cellwise constant functions and then solving that modified problem exactly for
a small interval of time. Then the new solution is again represented by piecewise constant
functions and the cycle is repeated. At first, this proposal was not taken very seriously by
practical scientists and engineers, who tended to see it as artificial, complicated and expen-
sive. Understanding and acceptance were eased by the explanation that it did not attempt to
find a discrete evolution operator, but instead worked by putting the data in a form to which
the exact operator could be applied, at least for short times.
Godunov [14] has written an interesting account of the circumstances under which this
work was done. He and his colleagues seem to have attached great importance to the inter-
pretation given above, and they based their approach to two-dimensional problems on taking
it as literally as possible, even though this forced them to confront a considerable difficulty.
To quote…

The first question we faced in our attempts to generalize our method for two dimensions
was how to find a solution of the two-dimensional Riemann problem with arbitrary
initial conditions.… In order for a two-dimensional method to be absolutely similar
to the one-dimensional one we needed an analytical solution of the gas-dynamic
equations with four initial discontinuities coming together in a single point. Nat-
urally, we did not have these solutions at that time and they are still unknown (for
general initial conditions). At this point, a roguish suggestion was made which was
to use only the solutions of a classical Riemann problem involving only planar waves.
Thus, the interaction of the four cells with a common vertex was neglected altogether.
This removed a nice physical interpretation underlying the construction of the one-
dimensional scheme. Quite naturally, there were many arguments during the discussion
of this hardly justifiable suggestion. L. V. Brushlinkii… carried out the analytical
solution for an acoustic wave propagating in stationary media, which took into account
the interaction of the cells sharing a common vertex. His solution was implemented
in a scheme completely analogous to the one-dimensional one. To our surprise and
pleasure there were no significant differences. Afterwards, only the simpler model
was used.

The simpler version that is in general use today, based solely on one-dimensional Riemann
problems at cell edges, seems to have been a desperate measure resorted to only after the
apparently more physically complete model that included corner interactions had proved
both difficult and fruitless. Today, the irrelevance of the corner interactions is less surprising,
because we are well aware what the solutions to “corner Riemann problems” look like, and it
is hard to imagine that their addition, which would resemble Fig. 1, might improve anything.
The story is a little more complicated because several people, for example [6,21] have
found that including corner interactions DOES improve the results from an upwind code for
two-dimensional linear advection. But apparently that effect does not carry over to the Euler
equations. The seeming paradox is resolved by realizing that advection, in any number of
dimensions, is a scalar problem with a one-dimensional domain of dependence that is easily
identified and taken into consideration. Acoustic waves, on the other hand, behave one-

123
1096 J Sci Comput (2017) 73:1094–1114

Fig. 1 The flow that might result


from taking into account the
corner interactions in a first-order
finite-volume scheme. The
“four-way” solutions are taken
from [20]

dimensionally only in one dimension. In n dimensions they have an n-dimensional domain


of dependence that does not have a simple mapping onto the grid, and cannot be represented
by any finite number of one-dimensional interactions. It follows that advection is not a reliable
guide in more than one dimension if we actually wish to study acoustics, or systems such as
the Euler equations that contain acoustic behavior.
In many recent attempts to formulate a computational framework for the Euler equations,
it has been regarded as almost axiomatic that the discrete representation should be discontin-
uous.1 Any consistent prediction of the evolution then involves considering one-dimensional
waves propagating normal to the discontinuity, and therefore introduces Riemann solvers.
Almost inevitably, two discontinuities will intersect, and their interaction will either be incor-
porated at considerable expense with little effect, or will in most cases be discarded with the
admission that the physical model is incomplete.

2 What Else Could There Be?

“Multidimensional upwind” methods too numerous to cite have been published over the
years, including attempts of my own [32,33,37]. I give an incomplete summary of some of
this work in [35]. Most of this is still within the framework of discontinuous representations,
however, and therefore still relies on one-dimensional flow models. In one dimension, the
mathematical description of advective and acoustic phenomena is only distinguished by
nonlinearity, so that other very real and important differences between them cannot properly
be detected.
An attempt to avoid these limitations is to be found in the class of Residual Distribution, or
Fluctuation-splitting schemes, introduced in [34]. In the earliest versions, designed merely to
solve advection problems, the solution was represented in two dimensions by unique values
at the vertices of a triangular mesh, and the flux integral φ T calculated around the boundary

1 A significant exception is the class of methods deriving from the SUPG concept put forward by Hughes and
coworkers [16,17]. These are not considered here as being outside our main interest.

123
J Sci Comput (2017) 73:1094–1114 1097

of each element T . This is divided into three, or fewer, unequal parts. These are used to
update the vertex values of the element. Each vertex may receive updates from several of the
elements around it. The rules for making the divisions and updates take into account both
the flow direction and the element geometry. The flow is assumed to be one-dimensional,
but not necessarily in a direction related to the element. This produced results with much
crisper discontinuities than those from grid-aligned upwinding [8]. The generalization of
these methods, first to systems of equations, then to unsteady flows, and eventually to viscous
flows [1] has however followed a path that I find slightly unconvincing, despite its very real
success.
Most development has followed van der Weide and Deconinck [39], who proposed to
extend the analysis carried out for the scalar case to systems of equations by devising matrix-
valued formulas that were equivalent in the sense of reducing back correctly to the scalar case,
and possessed other useful properties. However, it has proved difficult to give these formulas
a clear independent interpretation other than as a superposition of plane wave solutions.
A different approach was taken by Mesaros [24] who divided the residual into parts that
represent elliptic and hyperbolic effects. Only the hyperbolic part of the residual is treated by
an extension of the scalar method. The elliptic part of the residual is reduced by a least squares
minimization that is direction-free. This method is the only ”upwind-inspired” method that I
know of to possess an accurate incompressible limit. Nishikawa et al. [26,30] extended it to
third-order and Nishikawa and van Leer [27] made it the basis for a transonic airfoil code that
achieved textbook multigrid. Unfortunately, there were difficulties going to three dimensions
or time-dependent flows. The approach to be described here is free from those limitations.
It takes a different approach to distinguishing varieties of multidimensional behavior, based
on distinguishing advective and non-advective behavior, rather than hyperbolic and elliptic
behavior.
In higher dimensions, almost all sets of conservation laws contain advective terms (v ·∇)u
that describe how some quantity is transferred from place to place. In the absence of other
terms, the current local value of that quantity would have been found, at a previous time,
somewhere along the particle path. This aspect of the problem deals with the propagation of
information by the moving medium. It is pure interpolation and because the most relevant
data is that closest to the particle path, upwinding is appropriate, and there is a clear basis
for limiting, because previous maximum and minimum values are available.
All other terms involve propagation of information through the medium. Acoustic waves
are the most obvious example. They are nonmaterial, and in the absence of advection the
domain of dependence is symmetrical. Upwinding is not necessary because data in every
direction is relevant. Because wave focussing may occur, there are no bounds on the behavior
of the solution, and therefore no clear and rigorous basis for limiting. Although the under-
lying motive for all “upwind” methods is purportedly to enforce the correct propagation of
informations, the current implementations achieve this only for one-dimensional problems.
The new methods described here are still finite-volume methods, but the solution is taken
to be continuous across cell boundaries, so that no Riemann problems arise. The fluxes
are instead independent variables obeying their own rules of evolution, and so the name
“Active Flux” method has been adopted. This almost automatically improves the accuracy
in comparison to a regular finite-volume scheme because each cell is associated with more
degrees of freedom. This is also true of discontinuous Galerkin methods, but to achieve any
particular order of accuracy, the Active Flux methods need less storage than discontinuous
Galerkin methods and do not rely on superconvergent behavior. It is in the evolution of
the fluxes that physics is accounted for. We will focus on third-order methods as being (in
my opinion) especially appropriate within the aerospace industry, although higher-order is

123
1098 J Sci Comput (2017) 73:1094–1114

certainly possible. But at third order the method remains rather simple, and is fully discrete
and explicit, allowing a time step up to the maximum permitted by the physics.
The present paper is the prelude to a series that will confront the issue of correctly prop-
agating different kinds of information in higher dimensions by dividing the Euler equations
into their advective and acoustic components. This is a form of operator splitting but at the
linear level it incurs no error because in fact these operators commute. It turns out that even for
nonlinear problems there are very few interaction terms, and these do not affect the accuracy
of the method.
The beginning of my most recent efforts to develop a method without discontinuities was
the investigation of a method due to Karabasov and Goloviznin [18] which also features
“active” fluxes. Together with my student Timothy Eymann, we derived a family of second-
order schemes of this form for one-dimensional linear advection, and came to realize [9] that
the optimal, third-order, member of the family was in fact the scheme derived in 1977 by van
Leer as part of his “new approach to numerical advection” [40] and denoted by him as Scheme
V. An appreciation of van Leer’s paper, with some additional one-dimensional schemes, has
been presented by Rider [31]. Several authors since van Leer have rediscovered this advection
scheme, apparently independently. See Popov and Ustyugov [29], and Akoh et al. [2]. A
large timestep version has been devised by Berthon et al. [4] and a semi-discrete form by
Zeng [42]. Akoh et al. [2] have made a version that they have applied to the two-dimensional
shallow water equations, although many details of their approach differ somewhat from ours.
For example, their reconstruction is not continuous and the method does not directly extend
to three dimensions. We employ a different extension of Scheme V, valid in any number of
dimensions, to compute the fluxes due to advection.
The application to wave equations is similar in principle to Scheme V, in that it is based on
the exact solution to a linear model problem. For the acoustic system this is an extension of
Poisson’s integral solution to the initial-value problem for the scalar wave equation [10]. For
the scalar wave equation itself, a related method was developed independently by Hagstrom
[3,15] and co-workers. The extension presented here to nonlinear acoustics in the presence
of vorticity was given by Fan and Roe [12], where it was shown to give far more isotropic
results than a DG method of the same formal accuracy. Nonlinear effects were included by
careful choice of local values for the coefficients in the equations.
This paper will focus on presenting motivational material and outlining the general
approach. Fuller details of the Active Flux method can be found in my students Ph.D. theses
[11,23], and will appear in papers that are under preparation.

3 Re-examining One-Dimensional Methods

3.1 Advection in One Dimension

van Leer’s Scheme V Let us stay for the moment with the one-dimensional advection equa-
tion, ∂t u + a∂x u = 0, which has certainly served as a testbed for the development of many
algorithms, including the Projection-Evolution-Reconstruction algorithms introduced by van
Leer in 1977 [40]. The general idea of these methods is to consider only data in the class that
can be described by some given set of basis functions b j (ξ ) where ξ = 2(x− Δx
x̄)
∈ [−1, 1] is a
local coordinate in each cell, and a j,k is the coefficient of b j in the cell k. The basis functions
are restricted by requiring that each of them vanishes for values of the argument outside the
range [−1, 1], and set

123
J Sci Comput (2017) 73:1094–1114 1099

u t
R
uj E’
u j-1/2 u j+1/2
F F’
E
ξ R
ξ E F ξ
a Δt
Fig. 2 Forming the flux for linear advection according to Scheme V by tracing characeristics. The flux through
the right boundary is 16 (u R + 4u F + u E )


k=k 
max j= jmax
u(ξ, 0) = a j,k b j (ξ ), (1)
k=1 j=1

which allows the exact solution to be written immediately as


k=k 
max j= jmax  
2at
u(ξ, t) = a j,k b j ξ− . (2)
Δx
k=1 j=1

The Godunov (or first-order upwind, or Scheme I) method results from choosing jmax ,
the number of basis functions, as one, with b1 (ξ ) = χ(ξ ) where χ(ξ ) is the characteristic
function of the cell, equal to 1.0 inside the cell and equal to 0.0 outside it. For this choice
a1,k is the average value in cell k. Another choice, called Scheme III, destined to become
the famous MUSCL scheme, involves taking two basis functions per cell. In addition to the
characteristic function, we include the gradient function

b2 (ξ ) = ξ χ(ξ ). (3)

Each of these choices leads to linear recontructions that are not C 0 continuous, and therefore
must be supplemented by a Riemann solver.
In [40] van Leer considered also the effect of adding quadratic variation to give a recon-
struction
3(1 − ξ 2 ) (1 + 3ξ )(1 − ξ ) (1 − 3ξ )(1 + ξ )
u j (ξ ) = ū j − u j− 1 − u j+ 1 , (4)
2 4 2 4 2

where ū is the mean value in the cell, and u L = u j−1/2 , u R = u j+1/2 are values at the left
and right cell boundaries. Because these boundary values u j+1/2 are shared by both of the
adjacent cells j, j + 1 this reconstruction is continuous. Therefore the required storage is two
units per cell, exactly the same as MUSCL. Because the boundary values are shared, however,
quadratic reconstruction is possible, and third-order accuracy can be achieved. He called this
Scheme V. Because the reconstruction is continuous, no Riemann solver is required. The
method is upwinded simply because the flux through each interface is found from the data
on the upwind side of it.
Note that the reconstruction makes no use of data outside the cell, so that the accuracy of
the reconstruction is independent of mesh geometry. This is a property that can and will be

123
1100 J Sci Comput (2017) 73:1094–1114

preserved in higher dimensions. Taking a > 0, and defining the Courant number ν = aΔtΔx ,
the flux leaving through the right-hand boundary is
 1  
¯
f j+ 1 = a u(ξ ) dξ = a ν(3 − 2ν)ū j − ν(1 − ν)u j− 1 + (1 − ν)2 u j+ 1 , (5)
2 1−ν 2 2

which is the average value in the shaded region of Fig. 2. It will be useful to observe that of
course exactly the same result is obtained by numerical integration (See again Fig. 2), first
calculating

3 1 1
f F  = f (1, 21 Δt) =a ν(2 − ν)ū j − ν(4 − 3ν)u j− 1 + (2 − ν)(2 − 3ν)u j+ 1 ,
2 4 2 4 2

(6)
 
f E  = f (1, Δt) =a 6ν(1 − ν)ū j − ν(2 − 3ν)u j− 1 + (1 − ν)(1 − 3ν)u j+ 1 , (7)
2 2

and then using these to integrate the flux by Simpsons Rule.


1
f¯j+ 1 = [ f R + 4 f F + f E ], (8)
2 6
because u F  = u F , etc. This version is helpful in the nonlinear case because then the exact
integrals become cumbersome.
Once we have the fluxes, we make a conventional finite-volume update
Δt  
ū n+1 = ū n
− f¯ 1 − f¯ 1 (9)
j
j
x j+ 1 − x j− 1 j+ 2 j− 2
2 2

to obtain a conservative scheme.2


It is exact for any quadratic data and therefore third-order
accurate. Because it depends only on the data in the upwind cell, the formula and its accuracy
are unchanged on irregular grids. It shares with the first-order upwind scheme the property
of remaining valid up to the outflow boundary. At inflow only the incoming flux needs to be
specified. Although this has the appearance of being a two-step method with an intermediate
time level, both levels are updated from the same initial data. The cost is essentially that of
a single-step method.
However, it is also useful to view it as having two components. The first is to update the
fluxes by a second-order upwind scheme, either by (5), or by (6), (7). The second component
is a central leapfrog scheme (9). This interpretation will be the key to multidimensional
generalizations.
The stability analysis can be carried out by writing the solution as a two-component vector
u = (ū j , u j+ 1 )T and the scheme is then
2

un+1 = GV un (10)
with

1 − 1 − e−iθ ν 2 (3 − 2 ν) − 1 − e−iθ ν (1 − ν) 1 − ν − ν e−iθ
GV = . (11)
6 ν (1 − ν) (1 − ν) (1 − 3 ν) − ν (2 − 3 ν) e−iθ
Because we now deal with a 2 × 2 system rather than a scalar equation this matrix has two
eigenvalues g1,2 . In many situations where more than one root arises, one of these roots would
2 By construction this fully discrete scheme is stable for ν ∈ [0, 1]. The transport step does not change the L
2
norm of the solution. The reconstruction step produces a quadratic solution inside each cell that has the same
edge values and integral as the transported solution, and it is easy to show that this process decreases the L 2
norm, on any mesh.

123
J Sci Comput (2017) 73:1094–1114 1101

be regarded as spurious, and was so described in [40]. This is the case with leapfrog schemes,
but here, the additional root exists because we have two independent degrees of freedom for
each cell, which means that we can represent frequencies beyond the Nyquist limit of θ = π.3
The two roots each represent legitimate solutions, in the separate ranges [0, π] and [π, 2π].
Thus the active flux schemes have twice the resolving power of regular schemes. Indeed
this is also true of Discontinuous Galerkin schemes. It is tedious but routine to confirm that
neither of the eigenvalues is outside the unit circle if ν ≤ 1.0. The amplification factor for a
signal proportional to eiθ is, for small θ
1
|g| = 1 − ν(1 − ν)(1 − ν + ν 2 )θ 4 + O(θ 6 ) (12)
72
and the largest absolute value for the coefficient of θ 4 in the range ν ∈ [0, 1] is 1/384. For
comparison, the standard third-order upwind finite-difference method has
1
|g| = 1 − ν(1 − ν 2 )(2 − ν)θ 4 + O(θ 6 ) (13)
24
and the corresponding coefficient has a maximum absolute value of 3/128.
The semi-discrete form of Scheme V has also been rediscovered [42]. It can be written as
    
ū j a 0 e−iθ − 1 ū j
∂t u = u j+ 1 . (14)
j+ 21 Δx 6 −4 − 2e−iθ 2

The matrix G V here is obtained as limν→0 ∂ν G.

The Discontinuous Galerkin method The dispersion analysis for discontinuous Galerkin
methods is given by Sherwin [38], Cheng and Shu [5], and by Kiridonova and Qin [19]. We
will not repeat this analysis but merely remind the reader of the main idea in the simplest
case. Within each cell j the solution is assumed to be piecewise linear, and may be written
in terms of two linear basis functions φ ± = 1 ± ξ
u j (ξ ) = (1 − ξ )u −j + (1 + ξ )u +j = φ − u −j + φ + u +j . (15)

The u ±j are values just inside cell j at ξ = ±1. They are not shared with the adjacent cells,
so there are again two degrees of freedom per cell. The governing equation is expressed in a
weak form as  ∞
ψ(ξ )((1 − ξ )u̇ −j + (1 + ξ )u̇ +j + a∂x u)dξ = 0, (16)
−∞
where dots denote time deriatives and ψ(ξ ) is one of two test functions. It can be shown that
the truncation error is minimised by taking the test functions equal to the basis functions,
so that ψ(ξ ) = φξ . After integrating the spatial derivative terms by parts we have the two
equations,
 1  1
− a −
− 2
(φ ) dξ u̇ j + φ + φ − dξ u̇ +j − f (−1) + u j + u +j = 0, (17)
−1 −1 2
 1  1
a −
φ + φ − dξ u̇ −j + (φ + )2 dξ u̇ +j + f (1) − u j + u +j = 0. (18)
−1 −1 2
where f (u) = au is the flux function. If these equations are integrated we find
 
1 − 1 + a +
Δx j u̇ + u̇ − f (−1) − u j + u −j = 0, (19)
3 6 2
3 The method is applicable even at boundaries, so no boundary conditions need be considered in the analysis.

123
1102 J Sci Comput (2017) 73:1094–1114

 
1 − 1 + a +
Δx j u̇ + u̇ j + f (1) + u j + u −j = 0. (20)
6 j 3 2
The interface numerical fluxes f (± 21 ) are assigned, as in the finite-volume method, by solving
a Riemann problem to obtain the upwind flux,4
f (−1) = au +j−1 , f (1) = au +j ,
leading to update formulas that can be expressed in the matrix form;-
 −   −
u a −3 −1 + 4e−iθ u
∂t = . (21)
u+ Δx 3 −1 − 2e−iθ u+
Although derived quite differently, this is actually identical to the semi-discrete version of
Scheme V described by Eq. 14. The matrix GDG appearing in the this equation has the same
trace and determinant as GV appearing in (14) so the matrices are similar, in the sense of
having the same eigenvalues. The two schemes are the same, apart from being expressed in
different bases. In fact
GDG = TGV T−1 (22)
 
2 −1
where T = is the matrix that transforms from the Scheme V basis of (mean value,
0 1
right state) to the DG basis of (left state,right state). As with Scheme V, the second eigenvalue
need not be thought of as spurious, but as giving results in the frequency range θ ∈ [π, 2π].

3.2 Comparing AF and DG1

We will compare the dispersion relationships of the semi-discrete schemes with those of
Scheme V at various Courant numbers. To compare the semi-discrete and fully-discrete
forms, we will assume perfect time integration for the semi-discrete form,
at 
u(t) = e Δx G u(0),
so that at t = Δx/a, equivalent to one timestep at a CFL number of unity, the amplification
factor is just G sd = e(g) , where g is the appropriate eigenvalue of G. A comparable measure
for fully-discrete schemes is to calculate an amplification factor
1
G f d = |g(ν)| ν , (23)
realizing that the number of timesteps required to reach t = Δx/a would be 1/ν. The
comparison is valid even if 1/ν is not an integer.
The results are compared in Figs. 3 and 4. There is a considerable reduction in damping
from use of the fully discrete scheme, especially in the range of frequencies θ ∈ [π/4, π/2].
The phase errors are also improved, especially at very high frequencies. The phase of Scheme
V is exact at ν = 0.5. This is a general result for schemes with upwind-biased stencils,
demonstrated by van Leer [40]. However, this property is lost in the semi-discrete version,
as also of course is exactness for ν = 1.0, even with exact time-stepping.
On this basis we can make an admittedly naive comparison between Scheme V and DG1,
which are both third-order methods.
1. Comparable results are obtained from Scheme V and DG1 when Scheme V uses half the
number of cells. This gives Scheme V the superior effectiveness by a factor of 2.

4 Other choices, such as the Rusanov flux, are possible but inferior.

123
J Sci Comput (2017) 73:1094–1114 1103

Fig. 3 The amplification factor


for Scheme V at Courant numbers
ν = 3/4 (longdash), 1/2 (dash),
and 1/4 (dashdot). The solid line
is the limiting case, CFL = 0,
which is also the amplification
factor for the DG1 method
|G|

arg(G)

Fig. 4 The phase speed for Scheme V at Courant numbers ν = 3/4 (longdash), 1/2 (dash, also exact), and
1/4 (dashdot). The solid line is the limiting result for CFL = 0, which is the phase error for the DG1 method
and all other semidiscrete forms

2. This allows time steps that are twice as large, increasing the effectiveness to a factor of
4.
3. Since RK3 requires three stages to complete a timestep, the factor becomes 12.
4. In two or three dimensions, the factor becomes respectively 24 and 48.
Without taking this analysis too seriously, it does suggest that the Active Flux method
has the potential to outperform DG1 by quite a large factor. Moreover, it appears that DG1
is only second-order accurate for the full Euler equations, whereas the Active Flux method
(which is the generalization of Scheme V) has so far proved to retain third order accuracy.
Any comparison should then perhaps be made with the substantially more expensive DG2.

DG with inexact time-stepping In practice the DG method will need to employ some time
integrator that is less than exact, and usually the time integration is performed by one of

123
1104 J Sci Comput (2017) 73:1094–1114

u t
R
uj E’
u j-1/2 u j+1/2
F F’
E
ξ R
ξ E F ξ
a Δt
Fig. 5 Forming the flux for Burgers’ equation by tracing characeristics. The flux through the right boundary
1 (u 2 + 4u 2 + u 2 )
is 12 R F E

the Runge–Kutta methods. The lowest-order time integrator that will preserve third-order
accuracy is of course RK3, which provides stability to the DG1 discretization only up to
ν = 0.409 so that the numbers in the list above should be multiplied by about 2.5. More
accurate timestepping, or the use of Strong-Stability-Preserving timemarching, will reduce
the available Courant numbers further.
Remark on Superconvergence It is well known that discontinuous Galerkin methods employ-
ing polynomial reconstructions of order p frequently exhibit an order of accuracy 2 p + 1
instead of the p + 1 that typifies finite-volume methods. For example, the DG1 method might
be expected to be second-order accuracy, but turns out to be third-order. The fact that it is
the semi-discrete limit of Scheme V goes a long way to explain this. Although Scheme V
only stores two degrees of freedom per element, each of them is actually shared between
two elements and alows the reconstruction of a quadratic rather than a linear function. If
DG1 is regarded in the same way, with the outflow value regarded as also being, for the pur-
pose of reconstruction, the inflow value of the downstream neighbor, a third-order quadratic
reconstruction can be performed. This conclusion applies at all orders of accuracy [36] and
the insight enables “superconvergent” Galerkin methods for linear advection to be derived
with any accuracy using continuous reconstruction. Observe, however, that this continu-
ous reconstruction is completely different from what is ususually described as continuous
Galerkin

3.3 Burgers’ Equation

A natural next step from linear advection is Burgers’ equation.

∂t u + u∂x u = 0. (24)

There is of course no unique way to extend the solution of a linear problem to a nonlinear
problem, but there is a very simple modification (see Fig. 5) that does extend to more general
problems. As for the linear advection equation, we first find the solution on the outflow
boundary x = x R by tracing back the characteristics from the quadrature points E  , F  . If
this characteristic issues from the point x = x R + x ∗ and carries the value u = u ∗ then

x ∗ = −τ u ∗ = −τ (u R + x ∗ ∂x u)

123
J Sci Comput (2017) 73:1094–1114 1105

Fig. 6 Evolution of a Gaussian profile according to Burgers’ equation

so that
uRτ
x∗ = − −u R τ (1 − τ ∂x u). (25)
1 + τ ∂x u
For linear advection, the characteristic originates at x ∗ = −aτ , so we are replacing the
constant advection speed a with a local speed u R (1 − τ ∂x u), where τ is either Δt or 21 Δt.
This simple first-order correction to the wavespeed carries over to the nonlinear wave equation
[12] and forms the basis for our treatment of the Euler equations.
Figure 6 shows a typical solution to Burgers’ equation using this method. The initial profile
is

u(x) = 0.05 + 0.95 exp −50(x − 0.5)2

which steepens into a shock at t = 0.173. The solution is found here at t = 0.150 just
before shock formation. Table 1 shows that a convergence rate close to 3.0 is achieved by
incorporating the nonlinear correction. Further from the shock, the convergence is much more
rapid. The treatment after a shockwave has formed will be part of a more detailed paper.

4 The Multidimensional Strategy

4.1 The Overall Plan

The purpose of this section is to set out a strategy for applying the ideas of Scheme V to
multidimensional nonlinear systems displaying both advective and acoustic behavior. This
will be explained at third order for two-dimensional problems on triangular grids. Non-
simplicial grids can also be handled but will not be discussed here. Arbitrary order is attainable
straighforwardly for linear problems, but nonlinear terms have to be handled with more care
for orders greater than three.

123
1106 J Sci Comput (2017) 73:1094–1114

Table 1 Accuracy of Burgers’ Solution

Grid No wavespeed correction With wavespeed correction

L2 error L2 order L2 error L2 order

80 5.4047E−04 – 1.2746E−04 –
160 5.1270E−04 0.08 3.0595E−05 2.06
320 1.2928E−04 1.99 5.2251E−06 2.55
640 3.2350E−05 2.00 7.6770E−07 2.77
1280 8.0974E−06 2.00 1.0352E−07 2.89
2560 2.0264E−06 2.00 1.3428E−08 2.95
5120 5.0689E−07 2.00 1.7308E−09 2.96

The essence of the method resides in applying appropriate numerics to appropriate physics,
which means treating advective and acoustic behavior differently. The data at time nΔt, that
is un , is updated to u∗ by considering pure inertial evolution with all pressure gradients turned
off. Then the partial solution at u∗ is updated to un+1 by considering only acoustic changes.
Both processes are treated by novel methods that combine low memory, large timesteps, and
high accuracy with continuous reconstructions.
The solution is represented in the usual finite-volume manner by giving the average values
of the conserved variables in every cell, together with some other information on the cell
boundaries. These will be used to compute the fluxes, although the fluxes themselves are not
suitable to store because they are usually ambiguous in specifying the state. We have used
the primitive variables at the vertices of simplex elements and at the midpoints of their edges
to define quadratic distributions. Such a representation reduces storage because much data is
now shared between elements, and therefore reduces the size of any linear algebra problems
that might arise5 One timestep consists of the following three stages;

1. The data on the boundary of each element is updated by some scheme that respects the
physics. This means that the advective terms are are applied first and upwinded and the
acoustic terms are applied to the partial update without upwinding. The two stages are
treated in more detail below.
2. All boundary data is converted to fluxes, and the cell average is updated by integrating
them. This raises the order of accuracy by one, compared to the boundary values.
3. The cell averages and boundary values are reconciled. We currently have two ways to
do this. One is to add a bubble function to the primitive variables, to produce agree-
ment between the cell averages when the primitive variables are integrated over the cell.
Another is to measure the discrepancy between those averages by making corrections to
the primitive variables.

For linear problems the first step is simple, because the advective and acoustic operators
commute. We can carry them out independently, in either order. In the nonlinear case, most
of the nonlinear terms can be absorbed into the coefficients of a linear problem, as with
Burgers’ equation, but a few terms need careful analysis. This will be described in a paper
that is being prepared.

5 This observation is in itself an argument against discontinuous reconstructions.

123
J Sci Comput (2017) 73:1094–1114 1107

4.2 A New Approach to Multidimensional Advection

Van Leers approach to advection has often been generalized to multidimensional linear
advection ∂t u + a · ∇u = 0. It is easily accomplished in principle by representing the data as
a sum of local basis functions within each cell. Each of these functions is translated by aΔt
and the new solution is projected onto the basis functions. The process has much in common
with the “remap” phase of a Lagrangian–Eulerian calculation [31] and is usually carried out
in the context of discontinuous reconstructions.
To deal with nonlinearity we take a different view, and regard the advective stage of the
Euler solution as one in which particles preserve their momentum until they collide. They
obey the pressureless Euler equations
∂t u + u · ∇u = 0, (26)
which is a generalization of Burgers’ equation to n dimensions and has a similar exact, but
implicit, solution, which is u = U(x − ut), for any arbitrary function U : Rn → Rn . so that
in a small neighborhood
x∗ = −tu∗ = τ (u + x∗ ∇u)
and
x∗ = (I − τ ∇u)−1 τ u (27)
This is the correct generalization of (25) However, (26) is not in conservation form, so
cannot be completely treated with the Active Flux method, although the second-order solution
u(0, τ ) = u(x∗ , 0) is good enough to update the point values. Then, the conservative form
of the momentum equations is employed for the final update
∂t (ρu) + ∇ · (ρu ⊗ u) = 0. (28)
This describes pure inertial motion provided that the continuity equation,
∂t ρ + ∇ · (ρu) = 0, (29)
is also satisfied, and this will require also the point values of ρ. These follow from continuity.
Along any pathtube we have ρ(t)S(t) = const, where S(t) is the area of a patch of fluid.
Each point of the patch moves with a constant speed u0 , so the mapping from location at
t = 0 to t = τ is
x(τ ) = x(0) + u(x(0))τ (30)
and  
∂x(τ ) 1 + τ ∂∂ux τ ∂∂u
yu
J= = . (31)
∂x(0) τ ∂∂vx 1 + τ ∂v
∂y

It follows that the density is given by


ρ(x∗ , 0)
ρ(0, τ ) = . (32)
det J(x∗ , 0)
This result is exact for the pressureless Euler equations until fluid paths collide, although
because we are using this result embedded into the full Euler equations, the pressure terms
there will prevent collision.
Numerical implementation is sketched in Fig. 7, which shows three triangular elements
and the piecewise parabolic distributions inside them. On each interface between triangles,
characteristics are traced back from the quadrature points needed for Simpsons Rule, in order

123
1108 J Sci Comput (2017) 73:1094–1114

F
n+1
n + 1/2
n
u y t y
x x

(a) (b)
Fig. 7 Particle paths are traced back from quadrature points on an interface extending in time. a Velocity
reconstruction, b streamline tracing

to locate the points x∗ . Care must be taken to avoid double counting when streamlines almost
coincide with an edge.

4.3 A New Approach to Multidimensional Wave Propagation

The scalar wave equation is ∂t φ = c2 ∇ 2 φ. for which the classical Poisson solution [7] gives
the general solution to the initial-value problem.
However, this is not the form in which wave behavior appears in the Euler equations.
Instead we find (at the linear level) the first-order system

∂t p + ∇ · u = 0 (33)
∂t u + ∇ p = 0 (34)

It is easy to show that the pressure always obeys the classical solution, and so can be found
from (35), but the velocity does not, unless the flow is irrotational. If the flow initially contains
vorticity, then we have simply ∂t ∇ × u = 0 and vorticity is preserved. Because the derivation
is quite lengthy, it is not given here, but the general solution to the initial-value problem is
 ct
p(x, t) =Mct p(x, 0) − ct Mct ∇ · u + t Mct ∇ 2 p(x, t) dt (35)
0
 ct
u(x, t) =Mct u(x, 0) − ct Mct ∇ p + t Mct ∇(∇ · u(x, t)) dt (36)
0

Just as with van Leer’s much earlier treatment of linear advection, a numerical method
follows from expressing the initial data in any finite-element basis and then evaluating these
integrals to predict the (active) fluxes on the boundaries of the control volumes. For piecewise
polynomial data the integrals have simple closed forms. Figure 8 shows how the residual at
each node is the sum of integrals over segments of circular discs in each neighboring element.
This can be efficiently coded as a loop over the elements.
Since a necessary condition for stability is that the numerical domain of dependence must
contain the physical domain of dependence, it follows that the time step must be restricted

123
J Sci Comput (2017) 73:1094–1114 1109

MR

(a) (b)
Fig. 8 Each node is updated by integration over the Mach disc enclosing it. a Cells contributing to node
update, b cells contributing to edge update

1.245 0.26
DG1 DG1
1.24 AF AF
0.25
1.235
Absolute Velocity

0.24
1.23
Density

1.225 0.23

1.22
0.22
1.215
0.21
1.21

1.205 0.2
1.3 1.35 1.4 1.45 1.3 1.35 1.4 1.45
r r
Fig. 9 Two snapshots of a spreading circular wave close to its peak. Red is DG1, blue is AF (Color figure
online)

so that no Mach disc escapes from the union of elements to which the corresponding node
belongs. It has been found in practice that time steps extremely close to that limit may be
used on unstructured grids.
Tests were carried out on the classical problem of an initial Gaussian pulse in pressure.
The nonlinear p-system was chosen, so that
∂t ρ + ∇ · u = 0, (37)
∂t u + ∇ p(ρ) = 0, (38)
with p(ρ) = ρ γ . The Active Flux method was found to offer much better resolution than
DG1 (which is still third-order for this problem). Figure 9 shows that the Active Flux method
does a much better job of capturing the peak value, and has only about a quarter of the vertical
scatter.
This approach may be compared with the evolution Galerkin method developed by
Lucakova-Medvidova and Morton [22], which also uses the wave equation as a jumping-
off point, but retains a discontinuous representation and makes an approximate integration
of the bicharacteristic equations. This only achieves second-order accuracy, on an enlarged
stencil.

123
1110 J Sci Comput (2017) 73:1094–1114

Fast Vortex after 1 period Fast Vortex after 1 period


10 10

10
10
Density Error

Density Error
10

10

10

AF 10
10 DG1
DG AF
DG1
DG
10 10
10 10 10 10 10 10 10
h work units
Fig. 10 Accuracy of the density in a translating vortex for four compressible flow methods, (left) as a function
of the degrees of freedom, and (right) as a function of the required work units

5 The Euler Equations

As already mentioned, the plan is to apply operator splitting to the advective and acoustic
aspects of the flow, despite the nonlinear nature of the problem. There are several options
for this, and so again the details are reserved for a future publication. However, it proves
possible to overcome the nonlinearities by absorbing them into the coefficients of a local
linear problem.
Tests have been carried out for the translating vortex problem, which was first suggested
by Chi-Wang Shu and has proved a most valuable resource in high-order code development.
The specific example used was taken from the High-Order Workshop [41].
Figures 10 and 11 show the errors in density and velocity from four different schemes.
Included for comparison are the results from a second-order MUSCL scheme. However,
this method was implemented only on a simple square mesh and does not have to deal with
geometrical terms. This code does feature a nonlinear limiter, although this was probably
not activated by this mildly nonlinear problem. The results for DG1 and DG2 were obtained
from a suite of programs written by Professor Chris Fidkowski. These codes were intended
for educational purposes to display results from a variety of methods. The results for Active
Flux come from research codes written by Doreen Fan and Brad Maeng as part of their work
toward their doctorates.
On the left, the methods are compared for given degrees of freedom, by plotting against the
effective mesh size h = (d.o.f.)−1/2 . The MUSCL and DG1 results are clearly second-order
accurate, and quite surprisingly similar. AF and DG2 are both clearly third-order accurate,
with some superiority attaching to DG2.
A comparison in terms of work units presented difficulty since none of the codes were
written with efficiency as a major objective. A crude handicapping method was adopted, in
which the degrees of freedom were multiplied by the number of updates needed to advance

123
J Sci Comput (2017) 73:1094–1114 1111

1
Fast Vortex after 1 period 1
Fast Vortex after 1 period
10 10

0 0
10 10
Velocity Error

Velocity Error
10 10

10 10

AF
10 DG1 10
AF
DG1

10 10
6 8
10 10 10 10 10 10 10
h
Fig. 11 Accuracy of the absolute velocity in a translating vortex for four compressible flow methods, (left)
as a function of the degrees of freedom, and (right) as a function of the required work units

through one unit of time. This factor was taken to be 1.5 for Active Flux, allowing for slightly
more complex operations, 2.0 for MUSCL to penalize for not dealing with geometry, 7.5 for
DG1, made up of three stages for RK3 and 0.4 for the Courant condition. For DG2, with six
stages for RK5 and a Courant condition of 0.235 [5], the factor was taken to be 21. A case
could be made that this is rather low because it does does not allow for the more complex
operations required.
In fact, there is much scope for argument about the fairness of these handicaps, but the
differences revealed are too large to be ignored. For this problem, and this range of grid
sizes, DG1 and DG2 turn to be quite comparable in accuracy, despite one being second order
and the other one third-order. But the Active Flux method is clearly superior to both. For
given work units the accuracy either in velocity or density is about one and a half orders of
magnitude better. For given accuracy the required work is lower by one order of magnitude.
Significantly, the Active Flux method could be run up to a CFL number of 1.0, as was the
case with our model problems.

6 Discussion

The striking success of high-resolution Finite-Volume methods, and discontinuous Galerkin


methods, both of which are based on discontinuous representations of the solution, has lead
the scientific computing community to think that such representations provide a necessary
key to success in the numerical solution of hyperbolic problems. This seems doubtful, because
it has long been realized that there are numerous theoretical objections, beginning with the
recognition by Godunov himself that carrying this philosophy to its logical conclusion was a
futile exercise. Other issues have been the poor representation of discontinuities lying oblique
to the grid, especially shear and contact discontinuities, and an excess of dissipation at very
low speeds. I attribute these shortcomings to the exclusive use of one-dimensional arguments
in attempting to inject physics into the algorithms.

123
1112 J Sci Comput (2017) 73:1094–1114

Some authors (myself included) have attempted to introduce directions that are determined
by the solution rather than by the grid, but I do not now find any of them convincing; they
compensate for bad physics rather than replacing it. The failure at low Mach numbers should
come as no surprise when we realize that one-dimensional unsteady incompressible flows
can only exist with constant velocity, and this makes the idea of a Riemann problem rather
absurd. Although numerous empirical extensions have been proposed, they do not compete
with codes specifically designed for incompressible flow.
The Active Flux method whose principles are described here avoids both of these issues by
not introducing unjustified discontinuities, and also by insisting that every type of behavior
in a multidimensional compressible flow shall be represented through an appropriate discrete
operator. That is, after all, precisely the philosophy behind upwinding as it is very success-
fully applied to one-dimensional flow. However, in higher dimensions there are a greater
number of distinct behaviors to be provided with distinct treatments. In consequence, many
of the techniques developed for compressible flow in recent decades do not apply, and new
techniques have had to be invented. So far, this has proved possible, and the new ideas have
proved to be simple and accurate. Much remains to be done, but comparison between the
new and the old is extremely encouraging.
In various tests, the method based on continuous reconstructions has proved more cost-
effective by least an order of magnitude. Good progress is being made at the moment on a
steady-state version, and on limiting methods. The principle challenge ahead is the exten-
sion to Navier–Stokes, and the intention is to adopt the hyperbolic versions of Navier–Stokes
recently put forward by Nishikawa [25] and by Peshkov and Romenski [28]. This will involve
considering, as in solid mechanics, both compressive and torsional waves, and it is encour-
aging that these will also commute, and may be treated independently.

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 Interna-
tional License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and
reproduction in any medium, provided you give appropriate credit to the original author(s) and the source,
provide a link to the Creative Commons license, and indicate if changes were made.

References
1. Abgrall, R.: A review of residual distribution schemes for hyperbolic and parabolic problems: the July
2010 state of the art. Commun. Comput. Phys. 11(04), 1043–1080 (2012)
2. Akoh, R., Ii, S., Xiao, F.: A multi-moment finite volume formulation for shallow water equations on
unstructured mesh. J. Comput. Phys. 229(12), 4567–4590 (2010)
3. Alpert, B., Greengard, L., Hagstrom, T.: An integral evolution formula for the wave equation. J. Comput.
Phys. 16(22), 536–543 (2000)
4. Berthon, C., Sarazin, C., Turpault, R.: Space-time generalized Riemann problem solvers of order k for
linear advection with unrestricted time step. J. Sci. Comput. 55(2), 268–308 (2013)
5. Cockburn, B., Shu, C.-W.: RungeKutta discontinuous Galerkin methods for convection-dominated prob-
lems. J. Sci. Comput. 16(3), 173–261 (2001)
6. Colella, P.: Multidimensional upwind methods for hyperbolic conservation laws. J. Comput. Phys. 87(1),
171–200 (1990)
7. Courant, R., Hilbert, D.: Methods of Mathematical Physics, vol. II. Interscience, New York (1962)
8. Deconinck, H., Paillere, H., Struijs, R., Roe, P.L.: Multidimensional upwind schemes based on fluctuation-
splitting for systems of conservation laws. Computational Mechanics 11(5–6), 323–340 (1993)
9. Eymann, T.A., Roe, P.L.: Active flux schemes. In: 49th AIAA Aerospace Sciences Meeting (2011)
10. Eymann, T.A., Roe, P.L.: Multidimensional active flux schemes. In: 21st AIAA Computational Fluid
Dynamics Conference (2013)

123
J Sci Comput (2017) 73:1094–1114 1113

11. Fan, D.: On the acoustic component of active flux schemes for nonlinear hyperbolic conservation laws.
Ph.D. thesis, Aerospace Engineering, University of Michigan (2017)
12. Fan, D., Roe, P.L.: Investigations of a new scheme for wave propagation. In: 22nd AIAA Computational
Fluid Dynamics Conference (2015)
13. Godunov, S.K.: A difference method for numerical calculation of discontinuous solutions of the equations
of hydrodynamics. Mat. Sb. 89(3), 271–306 (1959)
14. Godunov, S.K.: Reminiscences about difference schemes. J. Comput. Phys. 153(1), 6–25 (1999)
15. Hagstrom, T.: High-resolution difference methods with exact evolution for multidimensional waves. Appl.
Numer. Math. 93, 114–122 (2015)
16. Hughes, T.J., Mallet, M.: A new finite element formulation for computational fluid dynamics: III. The
generalized streamline operator for multidimensional advective–diffusive systems. Comput. Methods
Appl. Mech. Eng. 58, 305–328 (1986)
17. Hughes, T.J., Scovazzi, G., Tezduyar, T.E.: Stabilized methods for compressible flows. J. Sci. Comput.
43(3), 343–368 (2010)
18. Karabasov, S.A., Goloviznin, V.M.: Compact accurately boundary-adjusting high-resolution technique
for fluid dynamics. J. Comput. Phys. 228(19), 7426–7451 (2009)
19. Krivodonova, L., Qin, R.: An analysis of the spectrum of the discontinuous Galerkin method. Appl.
Numer. Math. 64, 1–18 (2013)
20. Lax, D., Liu, X.-D.: Solution of two-dimensional Riemann problems of gas dynamics by positive schemes.
SIAM J. Sci. Comput. 19(2), 319–340 (1998)
21. Levecque, R.J.: High-resolution conservative algorithms for advection in incompressible flow. SIAM J.
Numer. Anal. 33(2), 627–665 (1996)
22. Lukacova-Medvidova, M., Morton, K.W., Warnecke, G.: Finite volume evolution Galerkin methods for
hyperbolic systems. SIAM J. Sci. Comput. 26(1), 1–30 (2004)
23. Maeng, J.: (Brad), On the advective component of active flux schemes for nonlinear hyperbolic conser-
vation laws. Ph.D. thesis, Aerospace Engineering, University of Michigan (2017)
24. Mesaros, L. M.: Multi-dimensional splitting schemes for theEuler equations on unstructured grids. Ph.D.
thesis, Aerospace Engineering and Scientific Computing, University of Michigan (1995)
25. Nishikawa, H.: New-generation hyperbolic Navier–Stokes schemes: O (1/h) speed-up and accurate vis-
cous/heat fluxes. In: 20th AIAA Computational Fluid Dynamics Conference (2011)
26. Nishikawa, H., Rad, M., Roe, P.L.: A third-order fluctuation-splitting scheme that preserves potential
flow. In: AIAA Paper 2001–2595, 15th AIAA CFD Conference, Anaheim (2001)
27. Nishikawa, H., van Leer, B.: Optimal multigrid convergence by elliptic/hyperbolic splitting. J. Comput.
Phys. 190(1), 52–63 (2003)
28. Peshkov, I., Romenski, E.: A hyperbolic model for viscous Newtonian flows. Contin. Mech. Thermodyn.
28(1–2), 85–104 (2016)
29. Popov, M.V., Ustyugov, S.D.: Piecewise parabolic method on local stencil for gasdynamic simulations.
Comput. Math. Math. Phys. 47(12), 1970–1989 (2007)
30. Rad, M.: A residual distribution approach to the Euler equations that preserves potential flow. Ph. D.
thesis, Aerospace Engineering and Scientific Computing, University of Michigan (2001)
31. Rider, W.J.: Reconsidering remap methods. Int. J. Numer. Methods Fluids 76(9), 587–610 (2014)
32. Roe, P.L.: Fluctuations and signals-a framework for numerical evolution problems. In: Morton, K.W.,
Baines, M.J. (eds.) Numerical Methods for Fluid Dynamics, vol. 11. Oxford University Press, Oxford
(1982)
33. Roe, P.L.: Discrete models for the numerical analysis of time-dependent multidimensional gas dynamics.
J. Comput. Phys. 63(2), 458–476 (1986)
34. Roe, P.L.: Linear Advection Schemes on Triangular Meshes. Cranfield College of Aeronautics, Cranfield
(1987)
35. Roe, P.L.: Multidimensional upwinding. In: Abgrall, R., Shu, C.-W. (eds.) Handbook of Numerical Anal-
ysis, vol. 18, pp. 53–80. Elsevier, Amsterdam (2017)
36. Roe, P.L.: A simple explanation of superconvergence for discontinuous Galerkin solutions to u t +u x = 0.
Commun. Comput. Phys. 21(4), 905–912 (2017)
37. Roe, P.L., Beard, L.: An improved wave model for multidimensional upwinding of the Euler equations. In:
Thirteenth International Conference on Numerical Methods in Fluid Dynamics. Springer, Berlin (1993)
38. Sherwin, S.: Dispersion analysis of the continuous and discontinuous Galerkin formulations. In: Discon-
tinuous Galerkin Methods. Springer, Berlin (2000)
39. Van der Weide, E., Deconinck, H.: Positive matrix distribution schemes for hyperbolic systems. In:
Désidèri, J.A., Hirsch, C., Le Tallec, P., Pandolfi, M., Périaux, J. (eds.), ECCOMAS ’96, Computational
Fluid Dynamics, pp. 747–753. Wiley, New York (1996)

123
1114 J Sci Comput (2017) 73:1094–1114

40. Van Leer, B.: Towards the ultimate conservative difference scheme. IV. A new approach to numerical
convection. J. Comput. Phys. 23(3), 276–299 (1977)
41. Wang, Z.J., et al.: High order CFD methods: current status and perspective. Int. J. Numer. Methods Fluids
72(8), 811–845 (2013)
42. Zeng, X.: A high-order hybrid finite difference-finite volume approach with application to inviscid com-
pressible flow problems: a preliminary study. Comput. Fluids 98, 91–110 (2014)

123
View publication stats

You might also like