Implicit time integration is applied to the fully coupled RANS and k-ω equations, both for steady and unsteady computations.. In this context, the purpose of this chapter is to describe
Trang 3Vol 2 Adaptive High-Order Methods in Computational Fluid Dynamics
edited by Z J Wang (Iowa State University, USA)
Forthcoming
Vol 1 Computational Methods for Two-Phase Flows
by Peter D M Spelt (Imperial College London, UK), Stephen J Shaw (X'ian Jiaotong – University of Liverpool, Suzhou, China) & Hang Ding (University of California, Santa Barbara, USA)
Trang 4N E W J E R S E Y • L O N D O N • S I N G A P O R E • B E I J I N G • S H A N G H A I • H O N G K O N G • TA I P E I • C H E N N A I
World Scientific
Computational Fluid Dynamics
Editor
Z J Wang
Iowa State University, USA
Trang 5British Library Cataloguing-in-Publication Data
A catalogue record for this book is available from the British Library.
For photocopying of material in this volume, please pay a copying fee through the Copyright Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA In this case permission to photocopy is not required from the publisher.
ISBN-13 978-981-4313-18-6
ISBN-10 981-4313-18-1
All rights reserved This book, or parts thereof, may not be reproduced in any form or by any means, electronic or mechanical, including photocopying, recording or any information storage and retrieval system now known or to be invented, without written permission from the Publisher.
Copyright © 2011 by World Scientific Publishing Co Pte Ltd.
Printed in Singapore.
ADAPTIVE HIGH-ORDER METHODS IN COMPUTATIONAL FLUID DYNAMICS Advances in Computational Fluid Dynamics — Vol 2
Trang 6To My Family
Trang 8vii
experts on adaptive high-order methods in computational fluid dynamics (CFD) It covers several widely used, and still intensively researched methods, including the discontinuous Galerkin (DG), residual distribution, differential quadrature, k-exact finite volume, spectral volume/spectral difference, PNPM, and correction procedure via reconstruction methods The reasons for including such a wide coverage
of methods are to: (1) provide a single source of reference, (2) present a snapshot of the state-of-the-art, and (3) facilitate the observation of similarities and differences as well as pros and cons of these methods
In the present context, adaptive high-order methods refer to numerical methods that are capable of handling unstructured adaptive meshes with accuracy higher than second-order These methods are compact, scalable, capable of handling both complex physics and geometry, and suitable for modern parallel supercomputers and graphics processing units (GPUs) They are widely considered the next major breakthrough in CFD, and have already found applications in computational aeroacoustics, computational electromagnetics, vortex dominated flows, and large eddy simulation and direct numerical simulation of turbulent flows
A concerted effort was made to minimize overlaps among the chapters For example, the first 7 chapters describe different aspects of the DG methods, while the last 8 chapters are devoted to other high-order methods Main topics covered include innovative formulations, analyses, efficient solution and time marching algorithms, parallel implementation, turbulence modeling, discontinuity-capturing techniques, error estimates, hp-adaptations, and dynamic mesh techniques, etc
The book requires a graduate student level of understanding It should serve as an excellent source of information for CFD developers, educators, researchers, users, and students who are interested in the state- of-the-art and the remaining challenges in adaptive high-order methods
Trang 9without their hard work Finally, I’d like to thank Ying Zhou for producing the color cover graphic, and Varun Vikas for help with Latex
Z.J Wang Ames, Iowa June 30, 2011
Trang 10ix
Francesco Bassi, Lorenzo Botti, Alessandro Colombo, Antonio Ghidoni And Stefano Rebay
Higher-Order Finite-Element Discretizations in CFD
Laslo T Diosady and David L Darmofal
for Discontinuous Galerkin Methods
Tobias Leicht and Ralf Hartmann
with Time Accurate Local Time Stepping
Gregor J Gassner, Florian Hindenlang and Claus-Dieter Munz
for CFD
Jaime Peraire and Per-Olof Persson
Discontinuous Galerkin Methods
Jianxian Qiu
Schemes for Diffusion Revisited
Bram van Leer, Marcus Lo, Rita Gitik and Shohei Nomura
Trang 11Chapter 9: High-Order Finite-Volume Discretization of the 235
Euler Equations on Unstructured Meshes
Carl Ollivier-Gooch and Chris Michalak
Schemes for Hyperbolic Problems
Rémi Abgrall
(RBF-DQ) Method and Its Applications
Chang Shu
Discretizations
Chris Lacor and Kris Van den Abeele
Discretization of Steady Problems
Georg May and Antony Jameson
Trang 12CHAPTER 1 DISCONTINUOUS GALERKIN FOR TURBULENT
FLOWS
Dipartimento di Ingegneria Industriale,Universit`a degli studi di Bergamo,Viale Marconi 5, 24044 Dalmine (BG), Italy
∗francesco.bassi@unibg.it
†lorenzo.botti@unibg.it
Dipartimento di Ingegneria Meccanica e Industriale,
Universit`a degli Studi di Brescia,Via Branze 38, 25123 Brescia, Italy
¶stefano.rebay@ing.unibs.it
The purpose of this chapter is to present all the relevant features of a
high-order DG method developed over the years for the numerical
solu-tion of the RANS and k-ω equasolu-tions The method has been implemented
using orthogonal and hierarchical modal shape functions defined in the
real space The code can handle hybrid grids consisting of tetrahedra,
prisms, pyramids and hexahedra Implicit time integration is applied to
the fully coupled RANS and k-ω equations, both for steady and unsteady
computations A directional shock-capturing term, proportional to the
inviscid residual, is employed to control oscillations around shocks Most
of the numerical results presented in this chapter have been computed
within the EU-funded ADIGMA project to investigate the capability of
the method for aeronautical applications
1 Introduction
In recent years several high-order methods have been emerging as
practi-cal tools to go beyond the second-order accuracy of standard finite volume
1
Trang 13discretizations of PDEs on general unstructured grids For aerospace
ap-plications this is of particular importance to further increase the impact of
Computational Fluid Dynamics (CFD) on the aerodynamic design of new
generation aircraft The Discontinuous Galerkin (DG) method, in
particu-lar, has been gaining popularity as one of the most promising approaches
to the accurate and robust numerical solution of ever more complex
physi-cal models and has attracted great efforts of many research groups into its
development
In this context, the purpose of this chapter is to describe several
devel-opments of the DG method implemented over the years in a fully parallel
DG code, named MIGALE, that we have used for the numerical solution
of the Euler, Navier-Stokes and the coupled RANS and k-ω turbulence
model equations These developments include: i) a proposal for adapting
the smooth-wall treatment of the variable ω to the degree of the polynomial
approximation, ii) the adoption of orthonormal and hierarchical modal
ba-sis functions defined in the real space for arbitrary shape elements, iii) a
shock-capturing technique based on the inviscid residual and applied in the
direction of the pressure gradient, iv) an implicit time integration technique
suited both for steady and unsteady problems
The capabilities of the present version of the code will be demonstrated
by computing several fairly complex problems taken from the suites of test
cases proposed within the EU-funded project ADIGMA In the conclusions
we will give a brief account of other recent implementations which are
already quite mature and will outline future directions of development
This section describes relevant implementation aspects of the DG
discretiza-tion applied to the coupled set of RANS and k-ω equadiscretiza-tions, including
approach and implicit time integration
Trang 14∂t(ρe0) +
∂
∂xj(ρujh0) = ∂
∂xj[uiτbij− qj] − τij
∂ui
∂xj+ β∗ρkee ω r, (3)
∂xj
+ τij
where the pressure, the turbulent and total stress tensors, the heat flux
vector and the eddy viscosity are given by
τij= 2µt
Sij−13
and turbulent Prandtl numbers and
Sij =12
is the mean strain-rate tensor The closure parameters α, α∗, β, β∗, σ, σ∗
Notice that the RANS and k-ω equations here employed are not in
in Eqs (3), (4), (5) Motivations for this choice have been discussed in
fulfills suitably defined “realizability” conditions, which set a lower bound
on eω in such equations This limitation substantially improves the stability
and robustness of turbulent flow computations because there is numerical
Trang 15evidence that too small, though positive, values of ω = eω e can lead to
sudden breakdown of computations
Realizability conditions, which guarantee that the turbulence model
pre-dicts positive normal turbulent stresses and satisfies the Schwarz inequality
for shear turbulent stresses, lead to the following inequalities
eω e
α∗ − 3
Sii−13
realizable turbulent stresses is given by
eω e r0
The solution of Eq (14) is trivial for the high-Reynolds number k-ω
Trang 162.1.1 Surface boundary condition for eω
2
where y is the local coordinate normal to the wall Eq (18) admits the
following near wall solutions
e
ω = log
6νβ
βeω e w
where eωwis the value of eω at the wall (y = 0) Of course, these solutions are
nothing but the logarithm of the viscous sublayer solutions for ω reported
appropriate solution for smooth walls, whereas non-singular solutions are
effects of surface roughness through surface boundary conditions
It has been shown in Ref 1 that singular and non-singular solutions for
smooth wall solution is recovered when the surface roughness tends to zero
In the rough-wall method surface values of ω can be simply set by means
of the correlation
ωw= Sr
u2 τ
the grid density or the degree of polynomial approximation should be high
enough to provide accurate solutions
On the other hand, the implementation of the smooth wall boundary
condition for ω requires special care in the numerical treatment of the
Trang 17Relying on the rough-wall method, the approach recommended by
Wilcox is simply to skip the issue of the numerical treatment of the
sin-gularity by replacing the perfectly smooth surface with an hydraulically
smooth surface In this so-called “slightly-rough-wall” boundary condition,
r < 5, i.e., it should ensure that the surface is hydraulically smooth with
roughness peaks lying within the viscous sublayer
The approach proposed by Menter consists in setting at the wall a finite
ωw= 106νw
by the factor 10, or, put another way, the analytical solution computed at
10
to find the equivalent slightly-rough-wall roughness implied by Menter’s
opti-mize the factor 10 of Menter’s formula by means of an accurate near-wall
numerical study of the ω solution and by comparing skin friction
distribu-tions of flat plate flows computed on differently refined grids The value of
the factor proposed by Hellsten is 1.25 instead of 10
In the framework of the DG method, an approach like that of Hellsten
was presented in Ref 2 where it was found that a good agreement between
experimental and numerical skin friction distributions of flat plate flows
6/β/50 As the solutions
keeping α not dependent on the polynomial degree of the solution did work
accurately However, as higher degree polynomials can follow closer and
closer the exact near wall distribution of ω, it seems reasonable to make α
dependent on the degree k of the polynomial approximation
Trang 18A possible approach in this direction has been outlined in Ref 5, where
with the value at the wall of the L2projection of the singular solution onto
the basis of the polynomial approximation
Here we propose an alternative way which consists of setting at the wall
the value eωk
around y = h, truncated to k terms, i.e.,
e
ωk
w= eωh− ∂ e∂yω
h
h1!+
∂2ωe
∂y2
h
h22! − · · ·
= eωh+
kXn=1
1
The finite values eωk
wat the wall can again be related to the exact solution
To actually apply Eq (25), h needs to be specified In the flat plate
computations presented below h has been set equal to the distance from
the wall of the centroid of the elements next to the wall As Eq (25) holds
for hydraulically smooth surfaces, we remark that the slightly-rough-wall
r < 5 If
polynomial solutions computed on relatively coarse grids targeted at
high-degree polynomial approximations
The boundary condition for ω has been tested on the flat plate flow
wall The Figure 1 displays the skin-friction distribution along the plate and
the profiles of u velocity component and of turbulence quantities at x/L =
between DG results and “average” experimental data is an effect produced
by the high-Reynolds number k-ω model here employed, that disappears
Trang 19Wieghardt law of the wall
300
law of the wall P 3
law of the wall P 4
law of the wall P 5
law of the wall P 6
DG - P 3
DG - P 4
DG - P 5
DG - P 6
using the modified coefficients of the low-Reynolds number version of the
model
The governing equations can be written in compact form as
∂u
source terms, Fc, Fv∈ RM⊗ RN denote the inviscid and viscous flux
func-tions, respectively, and N is the space dimension
Trang 20The weak form of Eq (27) reads
Z
+ZΩφs(u, ∇u) dx = 0, (28)where φ denotes any arbitrary, sufficiently smooth, test function and
Ω, consisting of a set of non-overlapping hybrid-type elements The
follow-ing space settfollow-ing of discontinuous piecewise polynomial functions for each
component uh i = uh 1, , uh M of the numerical solution uh is assumed:
of global degree at most k on the element K
The discontinuous approximation of the numerical solution requires
in-troducing a special treatment of the inviscid interface flux and of the viscous
flux For the former it is common practice to use suitably defined numerical
flux functions which ensure conservation and account for wave propagation
For the latter we employ the BR2 scheme, presented in Refs 8, 9 and
the-oretically analyzed in Refs 10, 11 (where it is referred to as BRMPS), to
obtain a consistent, stable and accurate discretization of the viscous flux
Accounting for these aspects, the DG formulation of problem (28) then
requires to find uh 1, , uh M ∈ Φh such that
Γ h
[[φh]] · bfu±h, (∇huh+ ηere([[uh]]))±
dσ+
Z
Ω h
φhs(uh, ∇huh+ r([[uh]])) dx = 0, (30)
average trace operators
[[q]]def= q+n++ q−n−, {·}def= (·)++ (·)−
where q denotes a generic scalar quantity and the average operator applies
to scalars and vector quantities By definition, [[q]] is a vector quantity
These definitions can be suitably extended to faces intersecting ∂Ω
ac-counting for the weak imposition of boundary conditions The local lifting
Trang 21operator re, which is assumed to act on the jumps of uhcomponentwise, is
defined as the solution of the following problem
Z
Ω h
φh· re(v) dx = −
Ze{φh} · v dσ, ∀φh∈ [Φh]N, v ∈L1(e)N
, (32)and the global lifting operator r is related to reby the equation
e∈E h
operators on the two sides of any edge e have support on the two elements
sharing the edge e Hence, the global lifting operator for any element
inde-pendently For the former we usually employ the Godunov flux or,
than the number of faces of the elements The BR2 viscous flux
discretiza-tion is as compact as possible because, for each element K, it only couples
the nearest neighbor elements This feature is obviously very attractive for
the implicit implementation of the method
2.2.1 Orthonormal and hierarchical basis functions
The actual implementation of Eq (30) requires specifying the test and trial
several aspects of the DG discretization, namely, i) numerical efficiency,
ii) conditioning of the DG discrete operators, iii) capability of easily
han-dling complex-shape elements
Modal expansion bases defined in the physical space can be used for
irregular and polyhedral elements in a very straightforward manner
Fur-thermore, it is quite easy to construct hierarchical and orthonormal sets
of shape functions that overcome the ill-conditioning of element mass
ma-trices that becomes evident for high-degree polynomial approximations on
highly stretched and curved elements Complex applications presented in
the following are in fact based on this type of approximation The main
Trang 22drawback of such modal polynomial approximations is the cost of numerical
integration for elements with non-constant Jacobian mapping
h Defining on K a set {ϕi}, i = 1, , NDOF, of
is the number of degrees of freedom of complete polynomials of degree k
Simple choices for {ϕi}, such as the set of monomials {xlymzn: l+m+n ≤
k}, are not advisable in general and for the sake of improving stability and
efficiency a set of orthogonal polynomial basis functions is highly preferable
The procedure to produce a set of orthonormal basis functions on a generic
element K relies on the modified Gram-Schmidt (MGS) orthogonalization
algorithm The sole requirement of this procedure is the capability to
com-pute the integral of polynomial functions on the desired element shapes
Let us denote with {ϕi} and {bi}, i = 1, , NDOF, the set of orthonormal
basis functions we wish to construct and a starting set of linearly
inde-pendent basis functions defined on K, respectively The MGS procedure
with re-orthogonalization can be simply setup as shown in the following
Trang 23Line 2 indicates that orthogonalization is applied twice As reported in
Ref 13, this is enough to get a set of basis functions which are orthonormal
up to machine precision It can be shown that the above MGS algorithm
following system of equations
ϕi=
i−1Xj=1
orthogonal to the i − 1 already orthonormalized basis functions, whereas
created ϕi For i = 1, , NDOF these coefficients are given by
aij
aii = − (bi, ϕj)K, j = 1, , i − 1,1
aii
=
sZK
bi−Xi−1j=1(bi, ϕj)K
2dx,
rij = −aij
aii,
rii= 1
aii
hierarchical In fact, increasing the degree of polynomial approximation
changing the already existing orthonormal basis functions
As regards the starting set of basis functions {bi} for a generic element
K, we have found that a simple and effective choice is the set of monomials,
up to the prescribed degree, expressed in a local frame of reference having
its origin in the centroid of K and the coordinate axes coincident with the
principal axes of the element
Finally we remark that the MGS algorithm outlined above is also used
to compute the values of basis functions (and of their spatial derivatives, if
necessary) at any location other than those needed to compute the integrals
of lines 4 and 6 In such cases the symbols b and ϕ at lines 5, 7 and 8
denote values of starting and orthonormalized basis functions (or of their
derivatives) at the desired location
Trang 242.3 Time integration
The DG space discretization of Eq (30) results in the following system of
(nonlinear) ODEs in time
where U is the global vector of unknown degrees of freedom, M is a global
block diagonal matrix and R (U) is the vector of “residuals”, i.e., the vector
of nonlinear functions of U resulting from the integrals of the DG discretized
space differential operators in Eq (30) We remark that using the DOFs of
approxima-tion, then the matrix M represents the global block diagonal mass matrix,
which, using orthonormal basis functions, reduces to the identity matrix
as unknowns of the polynomial expansion of the solution, then the block
ZK
2.3.1 Linearly implicit Runge-Kutta schemes
Implicit time integration of Eq (36) can be efficiently performed by means
of linearly implicit Rosenbrock-type Runge-Kutta schemes The class of
methods here considered can be compactly written as
sXj=1
αijKj
− J
i−1Xj=1
γijKj, i = 1, , s,
(37)
where s is the number of stages, bi, αij, γij are real coefficients and J =
the Euler scheme and for the schemes proposed in Refs 14 and 15 are
summarized in Table 1
Trang 25Table 1 Coefficients of some linearly implicit Runge-Kutta schemes.
An implementation of Eq (37) that saves at each stage the cost of the
j=1γijKj can be written as follows
sXj=1
aijWj
i−1Xj=1
cijWj.The coefficients of the transformed scheme are given by
(m1, , ms) = (b1, , bs) Γ−1, (aij) = (αij) Γ−1,
C = diag γ−1, , γ−1
− Γ−1,where Γ−1 def= (γij)−1 denotes the inverse of the coefficient matrices of
The matrix-explicit or the matrix-free GMRES algorithm can be used
to actually solve Eq (38) at each time step In both cases system
precondi-tioning is required to make the convergence of the GMRES solver acceptable
in problems of practical interest The Jacobian matrix implemented in our
code has been derived analytically and takes full account of the dependence
of the residual on the unknown vector and on its derivatives, including the
Trang 26Table 2 Inverse of the (γ ij ) matrices of Table 1.
implicit treatment of the lifting operators and of the boundary conditions
Using a suitably accurate time integration scheme, this allows to employ
the implicit solver also for accurate unsteady computations
The choice of the time step can significantly affect both the efficiency
and the robustness of the method For steady computations we have
im-plemented the pseudo-transient continuation strategy with the local time
step given by
c + d,where
hK
SK,define convective and diffusive velocities and the reference dimension of
robust strategy to increase the CFL number as the residual decreases is not
an easy task, especially for turbulent computations The rule here proposed
is essentially the result of intensive numerical experimentation and aims at
Trang 27Fig 2 Streamlined body: Mach number contours of P 4
solution and residuals
(
x = min (xL 2, 1) if xL ∞ ≤ 1
number, the maximum CFL number of explicit schemes and the exponent
(usually ≤ 1) governing the growth rate of the CFL number, respectively
been found useful to prevent sudden breakdown of computations once the
CFL number has already reached quite high values For relatively simple
steady test cases, such as the flow around a streamlined body (Figure 2,
232969 DOFs), the implicit time integration combined with the above CFL
number evolution rule provides quadratic Newton convergence to machine
accuracy
The shock-capturing approach consists of adding to the DG discretized
equations an artificial viscosity term that aims at controlling the high-order
modes of the numerical solution within elements while preserving as much as
possible the spatial resolution of discontinuities The shock-capturing term
is local and active in every element, but the amount of artificial viscosity is
proportional to the (inviscid) residual of the DG space discretization and
thus it is almost negligible except than at locations of flow discontinuities
Trang 28The shock-capturing term added to Eq (30) reads
XK
ZK
p(u±h, uh) (∇hφh· b) (∇huh· b) dx, (40)with the shock sensor and the pressure gradient unit vector defined by
∂p(uh)
∂uh i
si(u±h), dp(uh) =
MXi=1
are actually the lifting of the interface jump in normal direction between
the numerical and internal inviscid flux components The further factor
fp(uh) in Eq (41) is a pressure sensor defined by
fp(uh) =|∇hp(uh)|
p(uh)
hKk
which improves the accuracy of solutions in regions with high but
other-wise smooth gradients and allows using the same value of the user-defined
parameter C (typically C = 0.2) for different degrees of polynomial
1 (∆x) 2+(∆y)1 2 +(∆z)1 2
where ∆x, ∆y and ∆z are the dimensions of the hexahedral enclosing K,
scaled in such a way that their product matches the volume of K The
shock-capturing technique outlined above is highly non-linear and
residu-als convergence of steady state solutions can be quite difficult, even
imple-menting a fully (linearized) implicit discretization of the shock-capturing
term (40) This is in fact the case for the solution of the transonic flow
Trang 29p T u k ln( ω ) CFL
solution and residuals convergence
80860 DOFs), shown in Figure 3, that requires quite a large number of
Newton iterations for convergence
3 Numerical Results
In this section we present the results of high-order DG solutions of several
complex turbulent flows of aeronautical interest All the computations
freestream conditions and the higher-order solutions from the lower-order
ones Solutions have been advanced in time by means of the linearly implicit
backward Euler method and the linear system (38) has been solved using
the default iterative solver available in PETSc, i.e., the restarted GMRES
algorithm preconditioned with the block Jacobi method with one block per
process, each of which is solved with ILU(0)
The flow around the three elements airfoil has been computed with a farfield
elements with curved, four-node edges, see Figure 4 The main difficulties
of this test are due to highly distorted elements shapes and to the flow
complexity of strongly interacting wakes, see Figure 5 Figure 6 displays
Trang 30Fig 4 L1T2: pressure and Mach contours of P 6
solution.
iterations and performance index units, which is a relative measure of CPU
for external flows that combines a simple geometry with complexities of
transonic flow, i.e., local supersonic flow, shocks, and turbulent
bound-ary layers separation The flow conditions are those of Test 2308, i.e.,
Trang 31Performance Index Unit
solution.
namic chord The grid consists of 215632 hexahedral elements with curved,
eight-node faces, shown in Figure 7 superimposed to the pressure contours
with 60 Krylov subspace vectors and 120 maximum iterations These
approximations All the computations have been run in parallel using 512
cores of the CINECA BCX/5120 cluster
In Figure 8 the pressure coefficient distributions of P2solution are
com-pared with the experimental data at seven sections along the span of the
wing The shock-capturing technique proves capable of providing accurate
Trang 32Fig 8 ONERA M6: pressure coefficient of P 2
solution (◦ 2156320 DOFs) compared with the experimental data (4).
resolution of the lambda shock structure all along the suction surface of the
wing and, unlike many results presented in the literature, the shocks can
still be clearly distinguished at section y/b = 0.8 Despite the quite coarse
sepa-ration near the wing tip, as shown in Figure 9 Table 3 reports the force
coefficients of P0→2solutions, showing that at least a one-degree higher P3
solution would be useful to assess the convergence of force coefficients
Trang 33Fig 9 ONERA M6: flow separation near the wing tip of P 2
solution.
approx-imation on a grid of 188928 hexahedral elements with curved, eight-node,
faces shown in Figure 11(a) The same Figure 11 displays the pressure
convergence history in terms of Newton iterations and performance index
algorithm with 60 Krylov subspace vectors and 120 maximum iterations
All the computations have been run in parallel using 512 cores of the DLR
Trang 34p u w k ln( ω ) CFL
Performance Index Units
10
p u w k ln( ω ) CFL
those computed by the TAU and FUN3D codes The FUN3D and TAU
solutions have been computed on a grid with 11459041 nodes and on an
employs 1889280 DOFs
The DLR-F6 wing-body transport configuration has been the object of
several wind-tunnel tests and computational studies, see Ref 19, and also
Trang 35solution (◦ 1889280 DOFs) compared with TAU (– – – 17053510 DOFs) and FUN3D (– · – 11459041 DOFs).
se-ries with the aim of assessing the state of the art of computational methods
as practical aerodynamic tools for aircraft force and moment prediction In
grids with 50618 and 404944 hexahedral elements with curved, eight-node
The parameters of the restarted GMRES solver have been set to 60 Krylov
subspace vectors and 120 maximum iterations for the coarse grid solutions,
and to 120 vectors and 480 iterations for the fine grid solutions Figure 16
shows the residuals convergence history of the coarse-grid solutions in terms
of Newton iterations and performance index units The coarse and fine grid
computations have been run in parallel using respectively 128 and 512 cores
Trang 36Table 4 DLR F6: force and pitching moment coefficients of DG solutions.
(a) coarse grid
of the DLR CASE cluster facility Figure 15 highlights the capability of the
P3solution to capture the detail of flow separation at the wing-root junction
coefficients computed on the coarse and fine grids There is a discrepancy
between the more accurate results on the two grids that needs to be further
investigated and no clear conclusion can be drawn about the asymptotic
values of the aerodynamic coefficients One issue could be the poor
geomet-rical approximation of surfaces when using only quadratic mappings for the
faces of very coarse meshes Finally, Figures 17, 18 and 19 give an overview
of the pressure coefficient and skin friction distributions of DG solutions
compared with reference results of the TAU and CFL3D codes taken from
Ref 18
solutions.
Trang 37Fig 15 DLR F6: wing-root juncture flow separation and turbulence intensity contours
10
p T u w k ln( ω ) CFL
Performance Index Unit
8
p T u w k ln( ω ) CFL
exper-imentally within the second international Vortex Flow Experiment
(VFE-2) The geometry here considered is the delta wing with large-radius
lead-ing edge, for which experimental pressure data are available in Ref 20
614770 tetrahedral and prismatic elements The prisms fill a few layers
of elements within boundary layers on the delta wing and on the sting
Trang 38The available grid points allowed to define only a linear mapping of
el-ement faces and this resulted in inaccurate pressure distributions on the
wing and sting surfaces All the computations have been run in parallel
using 512 cores of the CINECA HPC cluster facility The convergence of
residuals for this test case was quite difficult and, more importantly,
observed vortex breakdown on the wing suction surface This issue could
be related to the poor representation of the wing and sting surfaces and
needs to be further investigated Figure 20 shows the pressure coefficient
Figure 20 clearly highlight the better resolution of vortices provided by the
solution (◦ 1012360 DOFs) compared with TAU (—— 5102446 DOFs) and CFL3D (– – – 2256896 DOFs, – · – 7689088 DOFs, – ··
– 26224640 DOFs).
Trang 39solution (◦ 1012360 DOFs) compared with TAU (—— 5102446 DOFs) and CFL3D (– – – 2256896 DOFs, – · – 7689088 DOFs,
– ·· – 26224640 DOFs).
4 Final Remarks
In this chapter we have presented and demonstrated several well-tried
fea-tures of the DG code MIGALE, that has been developed over the years
for the numerical solution of the coupled RANS and k-ω turbulence model
equations
Open issues of the proposed DG method are mainly related to its
com-putational cost and this has motivated our most recent research efforts in
two directions
On the one hand, we have developed a spectral DG method, with a
cou-ple of choices for the sets of collocation and integration points, to improve
the computational efficiency and a p-multigrid strategy to reduce the RAM
required by the fully coupled implicit solver The p-multigrid algorithm has
been analyzed in Ref 21 and applied to the solution of the compressible
Trang 40solution (◦ 4049440 DOFs) compared with TAU (—— 5102446 DOFs) and CFL3D (– – – 2256896 DOFs, – · – 7689088 DOFs, – ··
– 26224640 DOFs).
Euler and Navier-Stokes equations in Refs 22 and 23 First applications
of the p-multigrid strategy to shockless turbulent flows around complex 3D
geometries have already provided encouraging results
On the other hand, we are working on exploiting the flexibility of the
modal DG discretization, with shape functions defined in the real space, to
improve the computational efficiency by means of agglomeration strategies
The agglomeration technique provides also the natural setting for the
devel-opment of h-multigrid solution strategies for high-order DG discretizations
First results of this research activity have already been reported in Ref 24
Finally, even if the shock-capturing approach turned out to be robust
and accurate, further research is needed to make its formulation fully
consis-tent with a residual-based artificial viscosity Moreover, the adverse impact
on the regularity of convergence of residuals needs to be further
investi-gated
... degree, expressed in a local frame of reference havingits origin in the centroid of K and the coordinate axes coincident with the
principal axes of the element
Finally we remark... to define only a linear mapping of
el-ement faces and this resulted in inaccurate pressure distributions on the
wing and sting surfaces All the computations have been run in parallel... a starting set of linearly
inde-pendent basis functions defined on K, respectively The MGS procedure
with re-orthogonalization can be simply setup as shown in the following