Numerical Methods For High-Speed Flows: Sergio Pirozzoli
Numerical Methods For High-Speed Flows: Sergio Pirozzoli
Sergio Pirozzoli
Università di Roma ‘La Sapienza’
Dipartimento di Meccanica e Aeronautica
Via Eudossiana 18, 00184 Roma, Italy;
e-mail: [email protected]
Abstract We review numerical methods for DNS and LES of turbulent compressible flow in the
presence of shock waves. Ideal numerical methods should be accurate and free from numerical
dissipation in smooth parts of the flow, and, at the same time, they must robustly capture shock
waves without significant Gibbs ringing, which may lead to nonlinear instability. Adapting to
these conflicting goals leads to the design of strongly nonlinear numerical schemes that depend
on the geometrical properties of the solution. For low-dissipation methods for smooth flows,
numerical stability can be based on physical conservation principles for kinetic energy and/or
entropy. Shock-capturing requires addition of artificial dissipation, in more or less explicit form,
as a surrogate of physical viscosity, to obtain non-oscillatory transitions. Methods suitable for
both smooth and shocked flows are discussed, and the potential for hybridization is highlighted.
Examples of the application of advanced algorithms to DNS/LES of turbulent, compressible
flows are presented.
CONTENTS
INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
FURTHER TOPICS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
Complex Geometries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
Time Integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
Discretization of Viscous Fluxes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
Annu. Rev. Fluid Mech. 2011
APPLICATIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
1 INTRODUCTION
Numerical simulation of high-speed flows has a long history, dating back to the
beginning of the computer era (von Neumann & Richtmyer, 1950; Courant et al.,
1952; Godunov, 1959). Several textbooks on numerical methods have appeared
over the years (Richtmyer & Morton, 1967; Hirsch, 1988; LeVecque, 1990; Laney,
1998; Toro, 2009). Important advancements have been made, but, as illustrated
by a recent comparative study (Johnsen et al., 2010) computational gasdynamics
has not yet converged to an ‘optimal’ computational strategy. The purpose of
this review is to check the status of the discipline, and systematize the enormous
amount of material produced over the years, a large part of which was reviewed by
Ekaterinaris (2005). Given the vastness of the subject, we mostly limit ourselves
to analyze the family of finite-difference (FD) schemes that are frequently used,
especially in the academic community, for direct numerical simulation (DNS)
and large-eddy simulation (LES) of compressible turbulent flows. Resolving the
range of scales present in these flows requires numerical schemes that are accurate,
robust, and efficient in terms of CPU requirements.
The reference physical model consists in the compressible Navier-Stokes equa-
tions for a calorically perfect gas, here written in integral form in Cartesian
coordinates
Z X3 Z
d
u dV + (fi − fiv ) ni dS = 0, (1)
dt V ∂V
i=1
where
ρ ρui 0
u = ρuj , fi = ρui uj + pδij , fiv = σij , j = 1, 2, 3,
ρE ρui H, σik uk − qi ,
(2)
are, respectively, the vector of conservative variables, and the vectors of the
convective and viscous fluxes in the i-th direction. Here ρ is the density, ui is
the velocity component in the i-th coordinate direction, p is the thermodynamic
pressure, E = e + ρu2 /2 is the total energy per unit mass, e = RT /(γ − 1) is
the internal energy per unit mass, H = E + p/ρ is the total enthalpy, R is the
gas constant, γ = cp /cv is the specific heat ratio, σij is the viscous stress tensor,
and qi is the heat flux vector. At high Mach numbers, the occurrence of very
strong shock waves and of severe viscous heating may bring to light additional
effects that are not incorporated in (1), such as chemical reactions and non-ideal
thermodynamic behavior (real-gas effects). These are not addressed in the present
review, and the interested reader should refer to specific publications (Bertin &
Cummings, 2006).
Under the assumption of smooth flow, the Navier-Stokes equations can be
equivalently cast in differential form
3 3
∂u X ∂fi X ∂f v
i
+ = . (3)
∂t ∂xi ∂xi
i=1 i=1
2
Numerical methods for high-speed flows 3
Equation (5) shows that the total kinetic energy only varies because of mo-
mentum flux through the boundary or to volumetric work of pressure forces
(which is zero for incompressible flow), whereas the convective terms do
not cause any net variation. This property has inspired numerical schemes
based on the attempt to enforce ‘kinetic energy preservation’ in the dis-
crete sense. Additional conservation laws can be derived from the Euler
equations for smooth flows, namely (Harten, 1983b)
Z Z
d
ρg(s) dV + ρg(s) ui ni dS = 0, (6)
dt V ∂V
The solutions of (1) exhibit tendency to form steep gradients (shock waves
and contact discontinuities), whose thickness is of the order of magnitude of the
mean-free-path, making shock ‘resolution’ with numerical methods infeasible in
most cases of practical interest. In the inviscid limit shock waves reduce to zero-
measure objects, across which the Rankine-Hugoniot jump conditions must be
satisfied. In order for the class of ‘weak’ solutions of the Euler equations to
coincide with that of the Navier-Stokes equations in the inviscid limit, additional
conditions (in the form of ‘entropy inequalities’) must be satisfied, to ensure
that the macroscopic effects of the small scale diffusion processes are properly
represented. For a thorough mathematical discussion of these issues the reader
may consult Lax (1973), Majda (1959), and LeVecque (1990).
To summarize, high-speed flows typically feature regions where the flow is
smooth, and the governing equations in their differential form hold, interspersed
by extremely thin regions, where the flow properties vary abruptly. Therefore,
it is not surprising that numerical methods for high-speed flows have specialized
into two classes, capable to deal with smooth flows and with shock waves, respec-
tively, and having very different properties. Indeed, it is known that standard
discretizations used for smooth flows cause (potentially dangerous) Gibbs oscilla-
tions in the presence of shock jumps, whereas typical methods used to regularize
shock calculations exhibit excessive ‘numerical’ viscosity. Schemes of the two fam-
ilies are illustrated in Sections 2 and 3, respectively. Tailored numerical methods
for high-speed flows should be able to put together the favourable properties of
the two basic families, which is frequently realized through their hybridization or
blending, as also explained in Section 3. The following discussion will mostly deal
with the discretization of the convective derivatives that appear in (3), this being
the main source of nonlinearity in the governing equations. As customary in the
computational gas-dynamics community, a method-of-lines approach is assumed,
whereby time integration is handled separately, using ODE dedicated methods.
Time integration, as well as discretization of the viscous fluxes and specification
of boundary conditions, are addressed in Section 4.
Numerical methods for high-speed flows 5
evaluation of the nonlinear convective terms (Phillips, 1959). Such deficiency can
also be traced back to failure to discretely preserve quadratic invariants associ-
ated with the conservation equations (Lilly, 1965; Orlandi, 2000). While finite
viscosity may help to stabilize calculations, it is usually safer to revert to al-
ternative discretization techniques capable of ensuring stability in the inviscid
limit.
2.1 Upwinding
The ‘upwinding’ approach, commonly followed in the gas-dynamics community,
is based on the idea that solutions of the Euler equations propagate along char-
acteristics, and therefore a stable numerical method should also propagate its
information in the same characteristic direction (Moretti, 1979). The standard
practice (referred to as ‘flux vector splitting’ (Steger & Warming, 1981)) in the
FD framework is to split the flux function into ‘positive’ and ‘negative’ parts
f (u) = f + (u) + f − (u), associated, respectively, with non-negative and non-
positive propagation velocities, i.e. df + /du ≥ 0, df − /du ≤ 0, which are dis-
cretized by means of left- and right-biased approximations, respectively, thus
ensuring linear stability.
In the finite-volume (FV) framework upwinding is usually achieved through
the ‘flux difference splitting’ (or Godunov) approach. Considering a FV semi-
discretization of (4),
dv j 1
=− f (vj+1/2 ) − f (vj−1/2 ) , (14)
dt h
Rx
where v j (t) = 1/h x j+1/2 v(x, t)dx, is the spatial average of the approximate
j−1/2
solution over the cell Ij = (xj−1/2 , xj+1/2 ), a suitable reconstruction operator
is used to determine approximate left and right states at the cell interfaces,
±
vj+1/2 , and the interface flux f (vj+1/2 ) is replaced with the numerical flux re-
sulting from an exact (or approximate) Riemann solver (Roe, 1981; Toro, 2009),
formally fˆj+1/2 = R(vj+1/2
− +
, vj+1/2 ). For example, Roe’s approximate Riemann
solver (Roe, 1981) dictates
− + 1 − +
|a
j+1/2 |
+ −
R(vj+1/2 , vj+1/2 )= f (vj+1/2 ) + f (vj+1/2 ) − vj+1/2 − vj+1/2 ,
2 2
(15)
+ − + −
where aj+1/2 = (f (vj+1/2 ) − f (vj+1/2 ))/(vj+1/2 − vj+1/2 ) is a characteristic speed
associated with the intermediate state. In the original formulation of Godunov
(1959), piece-wise constant reconstructions were assumed, leading to the general
first-order flux
fˆj+1/2 = R(v j , v j+1 ). (16)
The extension to higher order of accuracy is achieved by replacing piece-wise
constant with piece-wise polynomial reconstructions (van Leer, 1979). In the
ADER (Arbitrary accuracy DERivative Riemann problem) approach (Toro et al.,
2001), the numerical flux is evaluated by solving the generalized Riemann problem
using a semi-analytical method. The approximate solution is given as a Taylor
time expansion at the cell interface position, thus allowing the construction of
one-step schemes with arbitrary accuracy in both space and time.
Numerical methods for high-speed flows 7
Upwinding has the main effect of damping the Fourier modes with highest
supported wavenumbers, with subsequent stabilizing effect on the numerical so-
lution. Upwind schemes have often been used for DNS of shock-free compressible
turbulence (Rai & Moin, 1993; Foysi et al., 2004; Pirozzoli et al., 2008), with a
good degree of success. However, as pointed out by Park et al. (2004), the numer-
ical dissipation introduced by upwinding can be harmful for LES, where proper
resolution of marginally resolved wavenumbers is crucial, as it may hamper the
effect of sub-grid scale models.
2.2 Filtering
Filtering the computed solution is a commonly used practice to cure nonlinear
instabilities of central schemes, while retaining high-order accuracy. High-order
filters were introduced by Lele (1992), and used by Visbal & Gaitonde (1999)
and Visbal & Gaitonde (2002) for the solution of conservation laws on stretched,
curvilinear, and deforming meshes. The latter authors considered compact filters
of the form
L
X dℓ
αv̂j−1 + v̂j + αv̂j+1 = (vj+ℓ + vj−ℓ ) , (17)
2
ℓ=0
where the coefficients dℓ at the r.h.s. are constrained to be functions of the free
parameter α through Taylor- and Fourier-series analyses, so as to achieve high-
order accuracy, and selectively damp the supported Fourier modes with highest
wavenumbers. A selective filtering strategy was developed by Bogey & Bailly
(2004) based on optimized explicit filters of the type (17) with α = 0, L = 4, 5, 6.
Those authors also suggested possible use of the selective filtering to model the
effects of the unresolved (sub-grid) flow scales, replacing classical eddy-viscosity
models. As for upwinding, filtering can eliminate spurious instabilities, but it
implies artificial energy draining that adds to the physical energy draining caused
by molecular viscosity. More accurate control over diffusive effects is achieved by
reverting to numerical methods capable of conserving energy in the absence of
viscosity.
equation, H for the total energy equation. Discretization of the mass and mo-
mentum equations in the split form (18) implies kinetic energy preservation at the
semi-discrete level (Honein & Moin, 2004), provided the difference operators sat-
isfy the summation by parts (SBP) property (Strand, 1994) that automatically
holds in the case of periodic problems for central (both explicit and implicit)
approximations of any order (Mansour et al., 1979). From another viewpoint,
discretization of (19) guarantees minimization of the aliasing error (Blaisdell
et al., 1996). Additional robustness in the presence of strong density variations
is gained (Kennedy & Gruber, 2008) by expanding the convective derivatives in
the generalized form
∂ρui ϕ ∂ρui ϕ ∂ρϕ ∂ui ϕ ∂ρui
=α + β ui +ρ +ϕ
∂xi ∂xi ∂xi ∂xi ∂xi
∂ϕ ∂ui ∂ρ
+ (1 − α − 2β) ρui + ρϕ + ui ϕ . (20)
∂xi ∂xi ∂xi
This arrangement leads to semi-discrete energy conservation in the case α = β =
1/4. Ducros et al. (2000) showed that the skew-symmetric forms (18) and (19)
give rise to locally conservative schemes, when the derivative operators are dis-
cretized with explicit central formulas (with order up to six). Pirozzoli (2010)
has recently proved that this holds true for explicit central formulas of arbitrary
order of accuracy, and presented computationally inexpensive numerical fluxes
also for the skew-symmetric form (20). Apparently, compact derivative approxi-
mations applied to the skew-symmetric split form of the convective terms do not
lead to locally conservative schemes. For a formal mathematical formulation of
fully conservative, skew-symmetric splitting of the compressible Euler equations,
the reader may refer to Morinishi (2010).
benefits in terms of nonlinear stability) if the first derivatives in (22) are replaced
with SBP operators. The resulting schemes are referred to as split high-order en-
tropy conserving (SHOEC), since discrete entropy conservation in the sense (6)
follows (Tadmor, 1984). The performance of SHOEC schemes is apparently af-
fected by the choice of the splitting parameter β. Sandham et al. (2002) suggested
that a value of β ≈ 4 is appropriate for numerical simulation of compressible tur-
bulence, whereas β ≈ 2 is more appropriate for numerical simulation of isotropic
turbulence at zero viscosity (Honein & Moin, 2004). As all formulations of the
Euler equations that rely on the use of entropy variables, SHOEC schemes suf-
fer for reduced computational efficiency, owing to swapping from conservative to
entropy variables, that requires CPU-intensive transcendental operations.
Honein & Moin (2004) developed entropy-consistent schemes by applying the
skew-symmetric splitting to the Euler equations, upon replacement of the total
energy equation with the entropy equation. This approach preserves the integrals
of ρs and ρs2 . An equivalent alternative is to use the internal or total energy
equations, rearranged in such a way that the skew-symmetric form of the entropy
equation automatically follows.
Tadmor scheme.
The key to the success of WENO reconstructions is the proper selection of the
‘linear’ weights dℓ , so as to achieve maximum order of accuracy (2L), and the
design of the ‘nonlinear’ weights ωℓ , such that if the stencil Sℓ contains a jump,
ωℓ ≈ 0 follows. Upwinding is achieved by suitable biasing in the reconstruction
procedure. Referring to the ‘positive’ part of the flux, in the original approach of
Jiang & Shu (1996) a left-biased reconstruction was obtained discarding the right-
most sub-stencil (i.e. dL = 0), thus reducing the overall stencil width (and the
formal order of accuracy) to 2L − 1. Weirs & Candler (1997) suggested the use of
a central stencil to reduce the numerical viscosity of WENO schemes, and biased
the reconstruction by re-defining the smoothness measurement for the right-most
sub-stencil, setting βL = max0≤ℓ≤L βℓ , thus preventing the selection of the fully
downwinded stencil (SL ). Formulas for the ‘negative’ split flux are easily obtained
by symmetry about xj+1/2 . Comprehensive compilations of WENO coefficients
cℓm , dℓ , γℓmn , can be found in Weirs & Candler (1997), Balsara & Shu (2000)
and Gerolymos et al. (2009).
12 Pirozzoli
Provable properties of WENO schemes are limited to the scalar case, and
include convergence for smooth solutions (Jiang & Shu, 1996), and existence and
stability of discrete shock profiles for the third-order version of the scheme (Jiang
& Yu, 1996). Nevertheless, WENO schemes are extremely robust in practice,
and have been applied successfully (both in their FD and FV versions) to a wide
number of physical problems. For an overview of the wide-ranging ramifications
of the WENO idea the reader may consult the review paper by Shu (2009).
Many investigators have tried to improve the properties of WENO schemes,
especially in terms of their behavior in smooth regions. The mainstream of re-
search (Weirs & Candler, 1997) has been focused on the attempt to modify the
linear weights and the polynomial reconstruction coefficients to improve the be-
havior in Fourier space and extend the range of well resolved wavenumbers, in the
spirit of the DRP approach. Follow-up studies of the ‘optimized’ WENO strat-
egy have been reported by Hill & Pullin (2004), Martı́n et al. (2006), and Taylor
et al. (2007). An alternative approach to conjugate the spectral-like resolution
properties of compact schemes with the shock-capturing properties of WENO was
undertaken by Deng & Zhang (2000), who developed weighted compact nonlin-
ear schemes (WCNS) by coupling WENO interpolation of conservative variables
to the intermediate nodes and compact conservative approximation of the flux
derivatives at the grid nodes. In the author’s experience, WENO schemes and
their variants are highly accurate, and quite robust. However, they are compu-
tationally expensive, mostly because of the large number floating point opera-
tions required for the evaluation of the smoothness measurements. Furthermore,
spectral-like resolution is never achieved owing to the noxious effect of the nonlin-
ear weights, unless the weights are locally frozen in smooth regions, as suggested
by Hill & Pullin (2004) and Taylor et al. (2007).
A visual impression of the performance of WENO schemes is gained from Fig-
ure 3, where the results of numerical simulations of the test case 13 of Lax &
Liu (1998) are reported. The solution of the two-dimensional Riemann problem
involves a complex shock diffraction pattern leading to the formation of a Mach
stem, and Kelvin-Helmholtz roll-up of the vertical slip-line. Figure 3 shows that
both the TVD and the WENO methods robustly capture the shock wave pattern,
without significant spurious oscillations. Significant differences are observed in
the predicted structure of the slip line, that exhibits finer scale roll-up for the
WENO schemes, and the formation of secondary vortex cores for the seventh-
order scheme.
tween the two schemes was established a-priori, and ENO was only activated
in the shock-normal direction, over twelve points upstream and twelve points
downstream of the mean shock location. Adams & Shariff (1996) first considered
a truly adaptive hybrid discretization, consisting of a baseline compact upwind
scheme coupled with a fifth-order ENO scheme. A simple switch, based on the
local gradient of the flux vector components, was used to mark critical cells, that
were subsequently padded with buffer cells on each side, to prevent the onset of
oscillations arising from the coupling of schemes with different properties. The
method was expanded by Pirozzoli (2002), who developed a fully conservative
formulation by hybridizing a fifth-order compact upwind numerical flux (of the
form (13)) with a seventh-order WENO flux, the switch being based on the local
density gradient. Figure 3d shows the results obtained with this method for the
Lax-Liu Riemann problem. Compared to baseline WENO schemes, the hybrid
formulation yields more accurate representation of smooth flow features, with
the formation of additional discrete vortices in the shear layer, but it also yields
some spurious wiggles near shocks. For this specific test case, the reduction of
computational cost with respect to standard seventh-order WENO is about 50%.
Further improvements to hybrid WENO schemes were introduced by Ren et al.
(2003) and Hill & Pullin (2004). In particular, the latter authors used a baseline
central DRP-like discretization (tuned for LES of compressible turbulence) and
designed a shock sensor based on the ratio of the largest to the smallest WENO
smoothness indicators λ = (maxℓ βℓ )/(minℓ βℓ + ε), that becomes large whenever
one (or more) WENO sub-stencils are crossed by discontinuities.
While hybrid schemes are now frequently used, it appears that the issues related
to coupling of the underlying schemes have not been thoroughly investigated.
A notable exception is the work of Larsson & Gustafsson (2008), who applied
Kreiss stability theory to the analysis of hybrid methods. Interestingly, it was
found that the coupling of two separately stable schemes may, in some instances,
give rise to an unstable system. However, it was also found that, if either scheme
is dissipative, the coupled system is strongly stable, thus endorsing the use of
upwind shock-capturing schemes around discontinuities.
The nonlinear filtering approach was introduced by Yee et al. (1999), who de-
signed a low-dissipative, shock-capturing algorithm based on SHOEC discretiza-
tion, augmented with the artificial dissipation of a TVD scheme. Referring to
the scalar conservation law, the following semi-discretization was proposed
dvj 1 ∗ ∗
= −Dfj − fj+1/2 − fj−1/2 , (27)
dt h
where the filter numerical flux is defined as
∗
fj+1/2 = k Ψj+1/2 dj+1/2 . (28)
approach over hybrid methods is that, in the case of explicit multi-stage time in-
tegration, the filter numerical flux can be applied at the end of the full time step,
rather than at each sub-step, with substantial saving in terms of CPU time (but
with some loss in terms of robustness). Ducros et al. (1999) modified the original
nonlinear filtering approach by adopting the artificial flux of the JST scheme (as
from Equation (30)), and replacing in Equation (28) the Harten switch with the
product of the Jameson sensor (32) and another shock indicator
(∇ · u)2
Θ= , 0 ≤ Θ ≤ 1, (29)
(∇ · u)2 + (∇ × u)2 + ǫ
frequently referred to as the Ducros sensor. Further modifications of the method
were introduced by Garnier et al. (2001), who replaced the TVD artificial flux
with the dissipative part of ENO and WENO schemes, and showed that the
Ducros sensor is capable to distinguish turbulent fluctuations from shocks better
than the Harten switch. The idea of adaptively filtering the numerical solution for
shock-capturing also underlies the work of Visbal & Gaitonde (2005) and Bogey
et al. (2009), who developed computational strategies based on high-accuracy
central discretization of spatial derivatives, supplemented with selective filtering
of the computed solution. Shock-capturing is achieved by locally decreasing the
filter bandwidth in critical regions.
A key ingredient in the formulation of hybrid and nonlinear filtering schemes
consists in the proper specification of the shock sensor. The Ducros sensor rep-
resents a simple and frequently used choice. Alternative simple shock sensors
include those proposed by Hill & Pullin (2004), Visbal & Gaitonde (2005), and
Larsson et al. (2007) (the latter being a modification of the Ducros sensor). More
elaborate sensors for shock waves and shears, based on multi-resolution wavelet
analysis were proposed by Sjögreen & Yee (2000) and Yee & Sjögreen (2007).
The performance of shock sensors in practical computations can be appreciated
from inspection of Figure 4, where we report results of application of some of the
indicators previously mentioned to the instantaneous flow field obtained from
DNS of transonic shock/boundary layer interaction (Pirozzoli et al., 2010). The
threshold values for the various sensors have been selected in such a way that
similar representation of the shock system is obtained. Note that the Ducros
sensor in its original formulation does not perform properly outside of the wall
layer (where ∇ × u ≈ 0), due to excessive sensitivity to dilatational fluctuations.
The sensor can be conveniently adapted to this case by setting ǫ = (u∞ /δ0 )2
in Equation (29), so that it is only activated when the local dilatation becomes
larger than a typical large-scale velocity gradient. Also note that, for evaluation
purposes, all other sensors are based on the pressure field. The data reported in
Figure 4 suggest that the modified Ducros sensor is capable to selectively isolate
‘genuine’ shocks, whereas other simple sensors also mark as critical zones smooth
regions populated by vortical structures.
constant), were added to the momentum and energy equations to spread shock
waves to a thickness comparable to the local mesh spacing. The artificial vis-
cosity idea was expanded by Jameson et al. (1981), who designed a high-order
non-oscillatory numerical flux by augmenting the numerical flux of a second-order
central scheme (fˆC ), with explicit diffusive fluxes resulting from discretization of
second- and fourth-derivative operators, as follows
fˆj+1/2 = fˆj+1/2
C
− dj+1/2 ,
h i
(2) (4)
dj+1/2 = aj+1/2 ǫj+1/2 (vj+1 − vj ) − ǫj+1/2 (vj+2 − 3vj+1 + 3vj − vj−1 ) , (30)
where in the case of the Euler equations the characteristic speed aj+1/2 is replaced
with the spectral radius of the Jacobian matrix of the inviscid flux at Roe’s
average state (Roe, 1981). The diffusion coefficients , defined as
(2) (4) (2)
ǫj+1/2 = k (2) ψj+1/2 , ǫj+1/2 = max 0, k (4) − ǫj+1/2 , (31)
Mani et al. (2009) replaced S with uk,k in (34), in the definition of crβ , and set
crβ to zero whenever uk,k > 0 (i.e. in expansion zones). Additional selectivity in
the definition of crβ was introduced by Bhagatwala & Lele (2009), in order that
it effectively becomes zero in the presence of resolved acoustic motions.
Further improvements of the artificial viscosity method have led to the inclusion
of the effects of strong temperature gradients (contact discontinuities), as well as
species diffusion in multi-component mixtures (Fiorina & Lele, 2007; Cook, 2007).
The extension of the method to generalized curvilinear coordinates was presented
by Kawai & Lele (2007).
∂ û ∂f (û)
+ = G1 + R, (35)
∂t ∂x
for the filtered variable û = G ⋆ u, where G is a (primary) low-pass filtering
operator. The term G1 accounts for the effect of the computed flow scales (i.e.
those represented on the computational grid) on the resolved ones (i.e. on the
filtered field), and is approximated as
∂f (û) ∂f (ũ)
G1 = − , (36)
∂x ∂x
where ũ = Q ⋆ û, Q being a regularized approximation of the exact deconvolution
operator G−1 . The effect of the interaction between the non-represented scales
and the resolved scales is further modeled through the relaxation term
4 FURTHER TOPICS
All methods presented so far are focused on the discretization of the spatial
derivatives that appear in the governing equations. Other issues, related to time
integration and specification of numerical boundary conditions, as well as the
extension to complex geometries, are briefly addressed in the present Section.
5 APPLICATIONS
Applications of the methods of computational gasdynamics to the analysis of
flow physics are innumerable. We limit ourselves to highlight the most recent
applications to high-fidelity simulation of flows involving the interaction of shock
waves with turbulence.
The simplest setting consists of the interaction of a (nominally) normal shock
wave with a field of isotropic turbulence. Canonical shock/turbulence interac-
tions were first investigated through DNS by Lee et al. (1997), using a hybrid
compact/ENO scheme. LES of the same problem was performed by Ducros
et al. (1999), by means of a characteristic-based nonlinear filtering scheme. The
problem was revisited by Larsson & Lele (2009), who carried out high-resolution
calculations using a hybrid central skew-symmetric/WENO discretization.
Substantial efforts in the last decades have been devoted to the analysis of
shock / boundary layer interactions (SBLI) (Délery & Marvin, 1986). The first
DNS study of SBLI was reported by Adams (2000), who investigated the flow
over a 18◦ ramp at free-stream Mach number M∞ = 3, using the hybrid com-
pact/ENO method of Adams & Shariff (1996). DNS of a 24◦ compression ramp
configuration at M∞ = 2.9, was performed by Wu & Martı́n (2007), using the
bandwidth-optimized WENO algorithm of Taylor et al. (2007). LES of the su-
personic ramp flow was carried out by Rizzetta et al. (2001), using the adaptive
filtering technique of Visbal & Gaitonde (2005), and by von Kaenel et al. (2004),
using the regularization method of Adams & Stolz (2002).
Another frequently used prototype SBLI consists of the reflection of an oblique
shock wave from a flat plate where a boundary layer is developing. The first LES
of impinging shock interaction was performed by Garnier et al. (2002), who used
a baseline central fourth-order discretization augmented with a nonlinear WENO
filter, whose local activation was controlled by the Ducros sensor. Pirozzoli &
Grasso (2006) performed a DNS study with flow conditions similar to Garnier
Numerical methods for high-speed flows 21
et al. (2002), using a seventh-order WENO scheme. LES of the impinging shock
interaction have also been performed by Touber & Sandham (2009). Their nu-
merical method relied on a baseline SHOEC scheme, and shock-capturing was
achieved through a TVD filter based controlled by the Ducros sensor. Sample
results of the calculations of Touber & Sandham (2009) are reported in Figure 5,
showing a comparison between the mean LES field (in terms of streamwise veloc-
ity and turbulent shear stress), and reference experimental data (Dupont et al.,
2008). Overall, the LES results are in very good agreement with experimental
PIV data, the most apparent difference being the size of the separation bubble.
However, the boundary layer thickening is very well captured, as well as the am-
plification of the Reynolds shear stress past the interaction zone, associated with
the shedding of vortices.
Shock/boundary layer interactions also occur under transonic conditions. The
first LES of transonic SBLI over a circular-arc bump was performed by Sandham
et al. (2003), using the SHOEC scheme with nonlinear TVD filtering. Pirozzoli
et al. (2010) have recently reported DNS results of transonic SBLI at M∞ =
1.3 over a flat plate using a hybrid approach, whereby smooth flow regions are
handled by means of conservative sixth-order central discretization of the skew-
symmetric split form (20), and shock waves are captured through a seventh-order
WENO scheme, the switch being based on the Ducros sensor. An illustrative
result of that calculation is reported in Figure 6, where the shock system is
captured through the modified Ducros sensor, and vortical structures through
iso-surfaces of the ‘swirling strength’ (Zhou et al., 1999). The figure highlights the
three-dimensional nature of the lambda-like shock pattern, whereby the vortical
structures in the incoming boundary layer cause the spanwise wrinkling of the
upstream compression fan. Numerous hairpin-shaped vortex loops, resembling
those found in incompressible boundary layer DNS, are observed both in the
upstream boundary layer and past the interacting shock.
The turbulent mixing resulting from the injection of an under-expanded sonic
jet in a supersonic cross-flow was studied by Kawai & Lele (2010) by means of
LES. Those authors used the artificial viscosity method in the version of Kawai &
Lele (2007), employing sixth-order compact approximations of the spatial deriva-
tives, coupled with eight-order low-pass filtering. The flow visualizations reported
in Figure 7 demonstrate the capability of the numerical method to capture the
front bow shock, the upstream separation shock, the barrel shock, and the Mach
stem, all without spurious wiggles, and at the same time to accurately resolve a
broad range of turbulence scales.
Low-dissipative shock-capturing methods have also been used by Hill et al.
(2006) to analyze Richtmyer-Meshkov instability with reshock. Those authors
used a DRP-like central approximation of the skew-symmetric split equations
in smooth regions, and switched to the tuned-WENO scheme of Hill & Pullin
(2004) near shock waves, the switch being controlled by the local ‘curvature’ of
the pressure and density field, in a fashion similar to the Jameson sensor.
Many other applications of low-dissipative shock-capturing algorithms are col-
lected in the review paper of Ekaterinaris (2005).
22 Pirozzoli
References
Adams NA. 2000. Direct simulation of the turbulent boundary layer along a
compression ramp at M = 3 and Reθ = 1685. J. Fluid Mech. 420:47–83
Adams NA, Shariff K. 1996. A high-resolution hybrid compact-ENO scheme for
shock-turbulence interaction problems. J. Comput. Phys. 127:27–51
Adams NA, Stolz S. 2002. A subgrid-scale deconvolution approach for shock-
capturing. J. Comput. Phys. 178:391–426
Arora M, Roe PL. 1997. On postshock oscillations due to shock capturing schemes
in unsteady flows. J. Comput. Phys. 130:25–40
Balsara D, Shu CW. 2000. Monotonicity preserving weighted essentially non-
oscillatory schemes with increasingly high order of accuracy. J. Comput. Phys.
160:405–452
Berland J, Bogey C, Bailly C. 2006. Low-dissipation and low-dispersion fourth-
order Runge-Kutta algorithm. Comp. & Fluids 35:1459–1463
24 Pirozzoli
Harten A. 1989. ENO schemes with subcell resolution. J. Comput. Phys. 83:148–
184
Harten A, Engquist B, Osher S, Chakravarthy SR. 1987. Uniformly high order
accurate essentially non-oscillatory schemes, III. J. Comput. Phys. 71:231–303
Harten A, Hyman JM, Lax PD. 1986. On finite-difference approximations and
entropy conditions for shocks. Comm. Pure Appl. Math. 29:297–322
Hill DJ, Pantano C, Pullin DI. 2006. Large-eddy simulation and multiscale mod-
elling of a Richtmyer-Meshkov instability with reshock. J. Fluid Mech. 557:29–
61
Hill DJ, Pullin DI. 2004. Hybrid tuned center-difference-WENO method for large
eddy simulations in the presence of strong shocks. J. Comput. Phys. 194:435–
450
Hirsch C. 1988. Numerical computation of internal and external flows. New York:
Wiley
Honein AE, Moin P. 2004. Higher entropy conservation and numerical stability
of compressible turbulence simulations. J. Comput. Phys. 201:531–545
Hu C, Shu CW. 1999. Weighted essentially non-oscillatory schemes on triangular
meshes. J. Comput. Phys. 150:97–127
Hu FQ, Hussaini MY, Manthey JL. 1996. Low-dissipation and low-dispersion
Runge-Kutta schemes for computational acoustics. J. Comput. Phys. 124:177–
191
Hu XY, Khoo BC, Adams NA, Huang FL. 2006. A conservative interface method
for compressible flows. J. Comput. Phys. 219:553–578
Jameson A. 2008a. Formulation of kinetic energy preserving conservative schemes
for gas dynamics and direct numerical simulation of one-dimensional viscous
compressible flow in a shock tube using entropy and kinetic energy preserving
schemes. J. Sci. Comp. 34:188–208
Jameson A. 2008b. The construction of discretely conservative finite volume
schemes that also globally conserve energy or entropy. J. Sci. Comp. 34:152–
187
Jameson A, Schmidt W, Turkel E. 1981. Numerical simulation of the Euler
equations by finite volume methods using Runge-Kutta time stepping schemes.
Paper 81-1259, NASA
Jiang GS, Shu CW. 1996. Efficient implementation of weighted ENO schemes.
J. Comput. Phys. 126:202–228
Jiang GS, Yu SH. 1996. Discrete shocks for finite difference approximations to
scalar conservation laws. SIAM J. Numer. Anal. 35:749–772
Johnsen E, Colonius T. 2006. Implementation of WENO schemes in compressible
multicomponent flow problems. J. Comput. Phys. 219:715–732
Johnsen E, Larsson J, Bhagatwala AV, Cabot WH, Moin P, et al. 2010. As-
sessment of high-resolution methods for numerical simulations of compressible
turbulence with shock waves. J. Comput. Phys. 229:1213–1237
Kawai S, Lele SK. 2007. Localized artificial diffusivity scheme for discontinuity
capturing on curvilinear meshes. J. Comput. Phys. 227:9498–9526
Numerical methods for high-speed flows 27
Kawai S, Lele SK. 2010. Large-eddy simulation of jet mixing in supersonic cross-
flows. AIAA J. Revision submitted
Kennedy CA, Carpenter MH, Lewis RM. 2000. Low-storage, explicit Runge-
Kutta schemes for the compressible Navier-Stokes equations. Appl. Numer.
Math. 35:177–219
Kennedy CA, Gruber A. 2008. Reduced aliasing formulations of the convective
terms within the Navier-Stokes equations. J. Comput. Phys. 227:1676–1700
Klein M, Sadiki A, Janicka J. 2003. A digital filter based generation of inflow
data for spatially developing direct numerical or large eddy simulations. J.
Comput. Phys. 186:652–665
Kok JC. 2009. A high-order low-dispersion symmetry-preserving finite-volume
method for compressible flow on curvilinear grids. J. Comput. Phys. 228:6811–
6832
Laney CB. 1998. Computational gasdynamics. Cambridge University Press
Larsson J, Gustafsson B. 2008. Stability criteria for hybrid difference methods.
J. Comput. Phys. 227:2886–2898
Larsson J, Lele SK. 2009. Direct numerical simulation of canonical
shock/turbulence interaction. Phys. Fluids 21:126101
Larsson J, Lele SK, Moin P. 2007. Effect of numerical dissipation on the predicted
spectra for compressible turbulence. Annu. Res. Briefs, Center for Turbulence
Research, Stanford University
Lax PD. 1973. Hyperbolic systems of conservation laws and the mathematical
theory of shock waves. In Regional Conference Series in Applied Mathematics.
SIAM
Lax PD, Liu XD. 1998. Solution of two-dimensional Riemann problems of gas
dynamics by positive schemes. SIAM J. Sci. Comput. 19:319–340
Lax PD, Wendroff B. 1960. Systems of conservation laws. Comm. Pure Appl.
Math. 13:217–237
Lee S, Lele SK, Moin P. 1997. Interaction of isotropic turbulence with shock
waves: effect of shock strength. J. Fluid Mech. 340:225–247
Lele SK. 1992. Compact finite difference schemes with spectral-like resolution.
J. Comput. Phys. 103:16
LeVecque R. 1990. Numerical methods for conservation laws. Basel: Birkhauser-
Verlag
Lilly DK. 1965. On the computational stability of numerical solutions of time-
dependent non-linear geophysical fluid dynamics problems. J. Comput. Phys.
93:11–26
Liu XD, Osher S, Chan T. 1994. Weighted essentially non-oscillatory schemes.
J. Comput. Phys. 115:200–212
Lui C, Lele SK. 2002. A numerical study of shock-associated noise. AIAA Paper
2002-2530, AIAA
Lund T, Wu X, Squires K. 1998. Generation of turbulent inflow data for spatially-
developing boundayr layer simulations. J. Fluid Mech. 140:233–258
28 Pirozzoli
von Kaenel R, Kleiser L, Adams NA, Vos JB. 2004. Large-eddy simulation of
shock-turbulence interaction. AIAA J. 42:2516–2528
von Neumann J, Richtmyer RD. 1950. A method for the calculation of hydrody-
namical shocks. J. Appl. Phys. 21:232–237
Wang ZJ. 2007. High-order methods for the Euler and Navier-Stokes equations
on unstructured grids. Progr. Aero. Sci. 43:1–41
Weirs VG, Candler GV. 1997. Optimization of weighted ENO schemes for DNS
of compressible turbulence. AIAA Paper 97-1940, AIAA
Wu M, Martı́n M. 2007. Direct numerical simulation of supersonic turbulent
boundary layer over a compression ramp. AIAA J. 45:879–889
Xu S, Martı́n MP. 2004. Assessment of inflow boundary conditions for compress-
ible turbulent boundary layers. Phys. Fluids 16:2623–2639
Yee HC. 1989. A class of high-resolution explicit and implicit shock-capturing
methods. Technical Memorandum 101088, NASA
Yee HC, Sandham ND, Djomehri MJ. 1999. Low-dissipative high-order shock-
capturing methods using characteristic-based filters. J. Comput. Phys.
150:199–238
Yee HC, Sjögreen B. 2007. Development of low dissipative high order fil-
ter schemes for multiscale Navier-Stokes/MHD systems. J. Comput. Phys.
225:910–934
Zhou J, Adrian RJ, Balachandar S, Kendall TM. 1999. Mechanisms for generating
coherent packets of hairpin vortices in channel flow. J. Fluid Mech. 387:353–396
Zingg DW, Lomax H, Jurgens HM. 1993. Optimized finite-difference schemes for
wave propagation. AIAA Paper 93-0459, AIAA
32 Pirozzoli
2 2
ρ′ /ρ0 /Mt 20
1.5 1.5
1 1
0.5 0.5
0 0
0 10 20 30 40 0 20 40 60 80
2 2
1.5 1.5
K/K0
1 1
0.5 0.5
0 0
0 10 20 30 40 0 20 40 60 80
0.2 0.2
S/cp /Mt 20
0 0
-0.2 -0.2
-0.4 -0.4
0 10 20 30 40 0 20 40 60 80
t/τ t/τ
D2 D6 D7 W7 S6 K6 H6 T2
Figure 3: Numerical simulation of test case 13 of Lax & Liu (1998) on a mesh
with hx = hy = 1/1200 (only the central part of the domain is shown). Time
integration is performed with a third-order SSP Runge-Kutta algorithm (the
Courant number was set to 0.6). Fourteen equally spaced density contours are
shown, from 0.6 to 2.3. (a) second-order TVD scheme with Van Leer limiter; (b)
fifth-order WENO scheme; (c) seventh-order WENO scheme; (d) hybrid com-
pact/WENO scheme (Pirozzoli, 2002). In the latter case, WENO is activated at
the intermediate node xj+1/2 (and two neighboring nodes left and right of it),
whenever |ρj+1 − ρj | ≥ 0.2.
34 Pirozzoli
25 25
20
(a) 20
(b)
15 15
y/δ0
10 10
5 5
0 0
15 20 25 30 35 40 15 20 25 30 35 40
25 25
20
(c) 20
(d)
15 15
y/δ0
10 10
5 5
0 0
15 20 25 30 35 40 15 20 25 30 35 40
x/δ0 x/δ0
Figure 4: Performance of shock sensors for DNS of transonic shock/boundary
layer interaction. Black lines denote pressure iso-lines (24 equally spaced levels,
1 ≤ p/p∞ ≤ 1.85). Red lines indicate iso-lines of shock sensors. (a) Ducros
sensor (Ducros et al., 1999); (b) Jameson sensor (Jameson et al., 1981); (c) Hill
& Pullin sensor (Hill & Pullin, 2004); (d) Visbal & Gaitonde sensor (Visbal &
Gaitonde, 2005). δ0 denotes the incoming boundary layer thickness.
Numerical methods for high-speed flows 35
12
10
y(mm)
12
10
y(mm)
x(mm)
Figure 5: LES of shock/boundary layer interaction (Touber & Sandham, 2009):
comparison with PIV experimental data (Dupont et al., 2008). Top: mean
streamwise velocity; bottom: Reynolds shear stress. The same contour levels
are used for PIV (filled contours) and LES (solid lines). The yellow/red lines
denote the mean separation line as from PIV (dashed) and LES (solid). Figure
courtesy of E. Touber and N. D. Sandham.
36 Pirozzoli
Figure 7: LES of jet in supersonic cross-stream (Kawai & Lele, 2010). Visualiza-
tions in streamwise/wall-normal plane are shown in the top row: density gradient
(left), and jet fluid (right). Visualizations in wall-parallel plane are shown in the
bottom row: streamwise velocity (left), and jet fluid (right). D is the jet diameter.
Figure courtesy of S. Kawai and S.K. Lele.