29.01.2015 Views

University of Patras - Nemertes

University of Patras - Nemertes

University of Patras - Nemertes

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>University</strong> <strong>of</strong> <strong>Patras</strong><br />

School <strong>of</strong> Engineering<br />

Department <strong>of</strong> Mechanical Engineering and<br />

Aeronautics<br />

A dissertation submitted to the<br />

Department <strong>of</strong> Mechanical Engineering and Aeronautics and the<br />

<strong>University</strong> <strong>of</strong> <strong>Patras</strong> in partial fulfillment <strong>of</strong> the requirements for the degree <strong>of</strong><br />

DOCTOR OF PHILOSOPHY<br />

in<br />

Mechanical Engineering<br />

High-order discontinuous Galerkin<br />

discretization for flows with strong moving<br />

shocks<br />

Author:<br />

Dipl.-Ing<br />

Konstantinos Kontzialis<br />

Supervisor:<br />

Pr<strong>of</strong>.<br />

John A. Ekaterinaris<br />

© Konstantinos Kontzialis. All rights reserved.<br />

<strong>Patras</strong> 2012


1<br />

Thesis committee:<br />

Pr<strong>of</strong>. John A. Ekaterinaris, Department <strong>of</strong> Mechanical Engineering and<br />

Aeronautics, <strong>University</strong> <strong>of</strong> <strong>Patras</strong><br />

Pr<strong>of</strong>. Yannis Kallinderis, Department <strong>of</strong> Mechanical Engineering and<br />

Aeronautics, <strong>University</strong> <strong>of</strong> <strong>Patras</strong><br />

Pr<strong>of</strong>. Kyriakos Giannakoglou, Department <strong>of</strong> Mechanical Engineering,<br />

National Technical <strong>University</strong> <strong>of</strong> Athens<br />

Examining committee:<br />

Pr<strong>of</strong>. John A. Ekaterinaris, Department <strong>of</strong> Mechanical Engineering and<br />

Aeronautics, <strong>University</strong> <strong>of</strong> <strong>Patras</strong><br />

Pr<strong>of</strong>. Yannis Kallinderis, Department <strong>of</strong> Mechanical Engineering and<br />

Aeronautics, <strong>University</strong> <strong>of</strong> <strong>Patras</strong><br />

Pr<strong>of</strong>. Kyriakos Giannakoglou, Department <strong>of</strong> Mechanical Engineering,<br />

National Technical <strong>University</strong> <strong>of</strong> Athens<br />

Pr<strong>of</strong>. Vasilis Kostopoulos, Department <strong>of</strong> Mechanical Engineering and<br />

Aeronautics, <strong>University</strong> <strong>of</strong> <strong>Patras</strong><br />

Pr<strong>of</strong>. Dimitris Saravanos, Department <strong>of</strong> Mechanical Engineering and<br />

Aeronautics, <strong>University</strong> <strong>of</strong> <strong>Patras</strong><br />

Pr<strong>of</strong>. Efstratios Gallopoulos, Department <strong>of</strong> Computer Engineering and<br />

Informatics, <strong>University</strong> <strong>of</strong> <strong>Patras</strong><br />

Pr<strong>of</strong>. John Tsamopoulos, Department <strong>of</strong> Chemical Engineering, <strong>University</strong> <strong>of</strong><br />

<strong>Patras</strong>


Acknowledgments<br />

The present thesis is the culmination <strong>of</strong> years <strong>of</strong> hard work. Many challenges and<br />

difficulties appeared during my research, which undoubtedly were the source <strong>of</strong><br />

inspiration and creation in many aspects <strong>of</strong> my PhD work.<br />

I am deeply thankful to my supervisor Pr<strong>of</strong>essor Yannis Ekaterinaris, a charismatic<br />

person, whose experience proved to be the source <strong>of</strong> guidelines and knowledge<br />

<strong>of</strong> unprecedented value. His patience, encouragement and spirit <strong>of</strong> giving, and most<br />

importantly his calm perseverance, were the key elements for the successful completion<br />

<strong>of</strong> my PhD research.<br />

I would like also to express my gratitude to Pr<strong>of</strong>essor Yannis Kallinderis who<br />

was the first person to initiate me in the wonderful field <strong>of</strong> CFD. His love for excellence<br />

taught me that ideas mature only through continuous and hard work.<br />

I express also my gratitude to the support <strong>of</strong> all the persons whose love and<br />

presence were <strong>of</strong> importance for a deeper understating and continuous effort.<br />

Lastly, I am deeply thankful to my very supportive family for the love and<br />

encouragement to keep on and succeed in my goals.<br />

2


Summary<br />

Supersonic flows over both simple and complex geometries involve features over a<br />

wide spectrum <strong>of</strong> spatial and temporal scales, whose resolution in a numerical solution<br />

is <strong>of</strong> significant importance for accurate predictions in engineering applications.<br />

While CFD has been greatly developed in the last 30 years, the desire and necessity<br />

to perform more complex, high fidelity simulations still remains.<br />

The present thesis has introduced two major innovations regarding the fidelity<br />

<strong>of</strong> numerical solutions <strong>of</strong> the compressible Navier-Stokes equations. The first one is<br />

the development <strong>of</strong> new a priori mesh quality measures for the Finite Volume (FV)<br />

method on mixed-type (quadrilateral/triangular) element meshes. Elementary types<br />

<strong>of</strong> mesh distortion were identified expressing grid distortion in terms <strong>of</strong> stretching,<br />

skewness, shearing and non-alignment <strong>of</strong> the mesh. Through a rigorous truncation<br />

error analysis, novel grid quality measures were derived by emphasizing on the direct<br />

relation between mesh distortion and the quality indicators. They were applied<br />

over several meshes and their ability was observed to identify faithfully irregularlyshaped<br />

small or large distortions in any direction. It was concluded that accuracy<br />

degradation occurs even for small mesh distortions and especially at mixed-type<br />

element mesh interfaces the formal order <strong>of</strong> the FV method is degraded no matter<br />

<strong>of</strong> the mesh geometry and local mesh size.<br />

Therefore, in the present work, the high-order Discontinuous Galerkin (DG)<br />

discretization <strong>of</strong> the compressible flow equations was adopted as a means <strong>of</strong> achieving<br />

and attaining high resolution <strong>of</strong> flow features on irregular mixed-type meshes<br />

for flows with strong moving shocks. During the course <strong>of</strong> the thesis a code was<br />

developed and named HoAc (standing for High Order Accuracy), which can perform<br />

via the domain decomposition method parallel p-adaptive computations for<br />

flows with strong shocks on mixed-type element meshes over arbitrary geometries<br />

at a predefined arbitrary order <strong>of</strong> accuracy. In HoAc in contrast to other DG developments,<br />

all the numerical operations are performed in the computational space,<br />

for all element types. This choice constitutes the key element for the ability to perform<br />

p-adaptive computations along with modal hierarchical basis for the solution<br />

expansion. The time marching <strong>of</strong> the DG discretized Navier-Stokes system is per-<br />

3


formed with the aid <strong>of</strong> explicit Runge-Kutta methods or with a matrix-free implicit<br />

approach.<br />

The second innovation <strong>of</strong> the present thesis, which is also based on the choice<br />

<strong>of</strong> implementing the DG method on the regular computational space, is the development<br />

<strong>of</strong> a new p-adaptive limiting procedure for shock capturing <strong>of</strong> the implemented<br />

DG discretization. The new limiting approach along with positivity preserving limiters<br />

is suitable for computations <strong>of</strong> high speed flows with strong shocks around<br />

complex geometries. The unified approach for p-adaptive limiting on mixed-type<br />

meshes is achieved by applying the limiters on the transformed canonical elements,<br />

and it is fully automated without the need <strong>of</strong> ad hoc specification <strong>of</strong> parameters as<br />

it has been done with standard limiting approaches and in the artificial dissipation<br />

method for shock capturing.<br />

Verification and validation studies have been performed, which prove the correctness<br />

<strong>of</strong> the implemented discretization method in cases where the linear elements<br />

are adequate for the tessellation <strong>of</strong> the computational domain both for subsonic and<br />

supersonic flows. At present HoAc can handle only linear elements since most grid<br />

generators do not provide meshes with curved elements.<br />

Furthermore, p-adaptive computations with the implemented DG method were<br />

performed for a number <strong>of</strong> standard test cases for shock capturing schemes to illustrate<br />

the outstanding performance <strong>of</strong> the proposed p-adaptive limiting approach.<br />

The obtained results are in excellent agreement with analytical solutions and with<br />

experimental data, proving the excellent efficiency <strong>of</strong> the developed shock capturing<br />

method for the DG discretization <strong>of</strong> the equations <strong>of</strong> gas dynamics.<br />

4


To my beloved family, to whom I owe everything.<br />

5


Nomenclature<br />

c k e(t) elemental degrees <strong>of</strong> freedom<br />

∆x e , ∆y e<br />

projections <strong>of</strong> the dual edges in the x, y directions<br />

∆x e,k , ∆y e,k<br />

grid metrics<br />

γ<br />

λ<br />

Λ<br />

Θ e<br />

J<br />

k<br />

adiabatic exponent<br />

thermal conductivity coefficient<br />

diagonal eigenvalue matrix<br />

auxiliary variable for gradient computation<br />

Jacobian matrix<br />

unit vector in an arbitrary direction<br />

L, R left and right eigenvectors<br />

n<br />

U<br />

u<br />

M kj<br />

R<br />

R e<br />

S k<br />

V k<br />

outward normal vector<br />

conservative variable vector<br />

contravariant velocity component<br />

mass matrix<br />

universal gas constant<br />

elemental residual<br />

line integral<br />

volume integral<br />

µ dynamic viscosity <strong>of</strong> the fluid<br />

6


7<br />

ω<br />

skewness angle<br />

e x xx, e x yy, . . . normalized error coefficients<br />

φ<br />

ρ<br />

τ<br />

θ<br />

shearing angle<br />

density<br />

two dimensional viscous stress tensor<br />

mesh rotation angle<br />

Ũ(x, t) solution over the computational domain<br />

Ũ e (x, t) FE interpolant<br />

Ω q st<br />

Ω e<br />

standard element configuration<br />

elemental space<br />

E(x, y) error in gradient computation<br />

b e k (x) elemental basis functions<br />

c<br />

c p<br />

c V<br />

d x , d y<br />

E<br />

e<br />

E x , E y<br />

local speed <strong>of</strong> sound<br />

specific heat at constant pressure<br />

specific heat at constant volume<br />

stretching factors<br />

total energy per unit volume<br />

internal energy per unit mass<br />

error in the computation <strong>of</strong> the x, y derivatives<br />

e x x, e x y, . . . error coefficients<br />

F i<br />

F v<br />

f x , f y<br />

H<br />

h<br />

inviscid flux tensor<br />

viscous flux tensor<br />

grid functions<br />

numerical flux<br />

(specific) enthalpy <strong>of</strong> the gas


8<br />

k<br />

L x , L y<br />

M<br />

Ma<br />

N<br />

p<br />

P r<br />

Boltzmann’s constant<br />

local element lengths<br />

Laplacian approximation<br />

Mach number<br />

order <strong>of</strong> approximation<br />

pressure<br />

Prandtl number<br />

Q, q mesh quality indices<br />

Q 1 , Q 2 , Q m , Q n<br />

number <strong>of</strong> quadrature points<br />

q x , q y<br />

Re<br />

s<br />

T<br />

heat fluxes<br />

Reynolds number<br />

entropy<br />

temperature<br />

u, v Cartesian velocity components<br />

u e<br />

v<br />

field value at the middle <strong>of</strong> each edge <strong>of</strong> the dual contour<br />

test or weighting function<br />

∇ h u, ∇u numerical and analytical value <strong>of</strong> the gradient<br />

S<br />

dual area


Acronyms<br />

TE Truncation Error<br />

FE Finite Element<br />

FD Finite Difference<br />

FV Finite Volume<br />

SD Spectral Difference<br />

SV Spectral Volume<br />

ENO Essentially non-oscillatory<br />

WENO Weighted essentially non-oscillatory<br />

DG Discontinuous Galerkin<br />

RK Runge-Kutta<br />

RKDG Runge-Kutta Discontinuous Galerkin<br />

TVB Total variation bounded<br />

SSP Strong stability preserving<br />

IC Initial conditions<br />

BC Boundary conditions<br />

EC Error coefficients<br />

CV Control Volume<br />

PETSc Portable Extensible Toolkit for Scientific Computing<br />

HoAc High Order Accuracy<br />

9


10<br />

GMRES Generalized Minimum Residual<br />

JFNK Jacobian Free Newton Krylov<br />

LDG Local Discontinuous Galerkin<br />

BO Baumann and Oden<br />

BR1 First Bassi-Rebay scheme<br />

BR2 Second Bassi-Rebay scheme<br />

NIPG Non-symmetric Interior Penalty Method<br />

DOF Degrees <strong>of</strong> freedom<br />

PDEs Partial differential equations<br />

ODEs Ordinary differential equations<br />

MPI Message Passage Interface<br />

GPUs Graphics Processor Units


Contents<br />

Contents 6<br />

1 Introduction 23<br />

1.1 Classical methods in CFD . . . . . . . . . . . . . . . . . . . . . . . . 27<br />

1.2 Why the DG method . . . . . . . . . . . . . . . . . . . . . . . . . . 31<br />

1.3 Thesis outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35<br />

2 Governing equations <strong>of</strong> gas dynamics 37<br />

2.1 Euler equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38<br />

2.1.1 Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38<br />

2.1.2 Polytropic ideal gas . . . . . . . . . . . . . . . . . . . . . . . . 39<br />

2.1.3 Entropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42<br />

2.1.4 Mathematical character <strong>of</strong> the Euler equations . . . . . . . . . 42<br />

2.2 Navier-Stokes equations . . . . . . . . . . . . . . . . . . . . . . . . . 44<br />

2.2.1 Mathematical character <strong>of</strong> the Navier-Stokes equations . . . . 45<br />

2.3 Non-dimensional form <strong>of</strong> the Navier-Stokes equations . . . . . . . . . 46<br />

2.4 Integral form <strong>of</strong> the compressible flow equations . . . . . . . . . . . . 48<br />

3 Mesh quality measures for the FV method 49<br />

3.1 Gradient computation in the FV method . . . . . . . . . . . . . . . . 52<br />

3.2 The general form <strong>of</strong> the truncation error for the node based FV method 53<br />

3.3 Consistency condition for the node based FV method . . . . . . . . . 56<br />

3.4 Verification <strong>of</strong> the error coefficients analytic expressions . . . . . . . . 59<br />

11


CONTENTS 12<br />

3.5 Elementary Types <strong>of</strong> Mesh Distortion . . . . . . . . . . . . . . . . . . 60<br />

3.6 Stretching . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62<br />

3.6.1 Skewness and Shearing . . . . . . . . . . . . . . . . . . . . . . 63<br />

3.6.2 Mesh rotation . . . . . . . . . . . . . . . . . . . . . . . . . . . 64<br />

3.6.3 Mixed-type element mesh interfaces . . . . . . . . . . . . . . . 65<br />

3.7 Direct relation between truncation error and mesh distortion . . . . . 66<br />

3.8 Definition <strong>of</strong> an appropriate index <strong>of</strong> mesh distortion . . . . . . . . . 68<br />

3.8.1 Normalized error coefficients . . . . . . . . . . . . . . . . . . . 68<br />

3.8.2 Mesh quality index . . . . . . . . . . . . . . . . . . . . . . . . 69<br />

3.8.3 Analytic expressions <strong>of</strong> the quality index Q for each type <strong>of</strong><br />

elementary mesh distortion . . . . . . . . . . . . . . . . . . . . 72<br />

3.8.4 Calibration <strong>of</strong> the mesh quality index Q . . . . . . . . . . . . 73<br />

3.9 Application <strong>of</strong> the mesh quality indices . . . . . . . . . . . . . . . . . 74<br />

3.9.1 Channel grids . . . . . . . . . . . . . . . . . . . . . . . . . . . 75<br />

3.9.2 Airfoil grids . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76<br />

3.9.3 Cylinder grids . . . . . . . . . . . . . . . . . . . . . . . . . . . 76<br />

4 Discontinuous Galerkin discretization 87<br />

4.1 Spatial Discontinuous Galerkin discretization . . . . . . . . . . . . . . 88<br />

4.1.1 Discontinuous weak formulation for the Euler equations . . . . 89<br />

4.1.2 The Numerical flux . . . . . . . . . . . . . . . . . . . . . . . . 92<br />

4.1.3 Basis functions and standard elemental configuration . . . . . 95<br />

4.1.4 Elemental operations . . . . . . . . . . . . . . . . . . . . . . . 100<br />

4.1.5 Selection <strong>of</strong> Gauss type numerical integration . . . . . . . . . 102<br />

4.1.6 Mapping between physical and computational space . . . . . . 103<br />

4.2 DG discretization for the Navier-Stokes Equations . . . . . . . . . . . 105<br />

4.2.1 Computation <strong>of</strong> the viscous terms in the DG method . . . . . 105<br />

4.2.2 The LDG method . . . . . . . . . . . . . . . . . . . . . . . . . 107<br />

4.3 The treatment <strong>of</strong> initial and boundary conditions . . . . . . . . . . . 110<br />

4.4 Wall boundary conditions . . . . . . . . . . . . . . . . . . . . . . . . 111


CONTENTS 13<br />

4.5 Far field boundary conditions . . . . . . . . . . . . . . . . . . . . . . 113<br />

4.6 Application <strong>of</strong> Initial Conditions . . . . . . . . . . . . . . . . . . . . . 115<br />

5 Time discretization 117<br />

5.1 Runge-Kutta methods . . . . . . . . . . . . . . . . . . . . . . . . . . 118<br />

5.2 Implicit time marching . . . . . . . . . . . . . . . . . . . . . . . . . . 121<br />

5.3 Newton’s method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122<br />

5.4 Krylov subspace methods . . . . . . . . . . . . . . . . . . . . . . . . 123<br />

5.5 Jacobian-free Newton-Krylov Method . . . . . . . . . . . . . . . . . . 124<br />

5.6 Preconditioning <strong>of</strong> the JFNK method . . . . . . . . . . . . . . . . . . 126<br />

6 Unified limiting 128<br />

6.0.1 Limiting for quadrilateral elements . . . . . . . . . . . . . . . 131<br />

6.0.2 Limiting for triangular elements . . . . . . . . . . . . . . . . . 134<br />

6.1 Positivity preserving limiters for the DG method . . . . . . . . . . . . 136<br />

7 Numerical results 139<br />

7.1 Solution output . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140<br />

7.2 Code Verification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141<br />

7.2.1 Convergence study for the Euler equations . . . . . . . . . . . 141<br />

7.2.2 Convergence study for the Navier-Stokes equations . . . . . . 143<br />

7.2.3 Inviscid flow over a cylinder . . . . . . . . . . . . . . . . . . . 144<br />

7.2.4 Standard Sod’s shock tube problem . . . . . . . . . . . . . . . 147<br />

7.2.5 Extreme Sod’s Riemann problem . . . . . . . . . . . . . . . . 149<br />

7.3 Supersonic flow over a cylinder . . . . . . . . . . . . . . . . . . . . . 151<br />

7.4 Double Mach Reflection <strong>of</strong> a strong shock . . . . . . . . . . . . . . . 157<br />

7.5 Inviscid flow at Mach number <strong>of</strong> 3 in a tunnel with a step . . . . . . . 163<br />

7.6 Diffraction <strong>of</strong> a strong shock over a backward facing step . . . . . . . 168<br />

7.7 Shardin’s problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178<br />

7.7.1 Strong Mach 5 shock impingement on a triangle . . . . . . . . 185<br />

7.8 Flat plate boundary layer . . . . . . . . . . . . . . . . . . . . . . . . 189


CONTENTS 14<br />

7.9 Unsteady viscous flow over tandem airfoils . . . . . . . . . . . . . . . 191<br />

8 Conclusions and Future Work 197


List <strong>of</strong> Figures<br />

2.1 Control Volume (CV) Ω. . . . . . . . . . . . . . . . . . . . . . . . . . 48<br />

3.1 Median dual surface for evaluating first-order derivatives at grid point<br />

0 via the FV method. . . . . . . . . . . . . . . . . . . . . . . . . . . . 52<br />

3.2 Mixed-type element channel mesh. . . . . . . . . . . . . . . . . . . . 60<br />

3.3 Distribution <strong>of</strong> the analytic (✷) and the numerical (✸) TE for the<br />

field u(x, y) = xy 2 on the mesh <strong>of</strong> Fig. 3.2 using the median dual. . . 61<br />

3.4 Difference between the analytic and the numerical TE for the field<br />

u(x, y) = xy 2 on the mesh <strong>of</strong> Fig. 3.2 using the centroid dual. . . . . 62<br />

3.5 Stretching for (a) structured and (b) unstructured meshes defined to<br />

facilitate comparison <strong>of</strong> accuracy degradation on them. . . . . . . . . 63<br />

3.6 Skewed mesh, depicting the deviation angle ω from the 180 ◦ angle<br />

between one <strong>of</strong> the two pairs <strong>of</strong> edges sharing point 0 (ω ∈ [0 ◦ , 90 ◦ ]).<br />

The lengths <strong>of</strong> the edges are equal. . . . . . . . . . . . . . . . . . . . 64<br />

3.7 Sheared mesh, depicting displacement <strong>of</strong> the edges causing deviation<br />

from orthogonality expressed by the angle φ (φ ∈ [0 ◦ , 90 ◦ ]). The<br />

lengths <strong>of</strong> the edges are equal. . . . . . . . . . . . . . . . . . . . . . . 64<br />

3.8 Rotated structured mesh depicting non alignment <strong>of</strong> the edges with<br />

the axes <strong>of</strong> the global system (∆x ≠ ∆y and θ ∈ [0 ◦ , 45 ◦ ]). . . . . . . 65<br />

3.9 Common mixed-type element mesh interfaces. . . . . . . . . . . . . . 65<br />

3.10 High aspect ratio quadrilaterals typical <strong>of</strong> a structured mesh in a<br />

boundary layer region. . . . . . . . . . . . . . . . . . . . . . . . . . . 71<br />

3.11 Mesh quality index Q x for the stretched unstructured mesh (◦) and<br />

the stretched structured mesh (□) vs. displacement factor d x (d y = 0). 73<br />

3.12 Mesh quality index Q x and its corresponding normalized error coefficients<br />

vs. the skewness angle ω. . . . . . . . . . . . . . . . . . . . . . 74<br />

15


LIST OF FIGURES 16<br />

3.13 Mesh quality index Q y for the mixed-type element mesh interface <strong>of</strong><br />

Fig. 3.9(a) vs. angle ψ 2 . . . . . . . . . . . . . . . . . . . . . . . . . . 75<br />

3.14 Locally stretched structured channel mesh: (a) mesh geometry, (b)<br />

index q and (c) index Q. . . . . . . . . . . . . . . . . . . . . . . . . . 78<br />

3.15 Locally stretched unstructured channel mesh: (a) mesh geometry, (b)<br />

index q and (c) index Q. . . . . . . . . . . . . . . . . . . . . . . . . . 79<br />

3.16 Mixed-type element channel mesh: (a) mesh geometry, (b) index q<br />

and (c) index Q. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80<br />

3.17 O–type structured mesh around a NACA 0012 airfoil: (a) mesh geometry,<br />

(b) index q and (c) index Q. . . . . . . . . . . . . . . . . . . 81<br />

3.18 Unstructured mesh around a NACA 0012 airfoil: (a) mesh geometry,<br />

(b) index q and (c) index Q. . . . . . . . . . . . . . . . . . . . . . . . 82<br />

3.19 Mixed-type element mesh around a RAE 2822 airfoil: (a) mesh geometry,<br />

(b) index q and (c) index Q. . . . . . . . . . . . . . . . . . . 83<br />

3.20 Structured mesh around a cylinder: (a) mesh geometry, (b) index q<br />

and (c) index Q. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84<br />

3.21 Unstructured mesh around a cylinder: (a) mesh geometry, (b) index<br />

q and (c) index Q. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85<br />

3.22 Mixed-type element mesh around a cylinder: (a) mesh geometry, (b)<br />

index q and (c) index Q. . . . . . . . . . . . . . . . . . . . . . . . . . 86<br />

4.1 Local reconstruction <strong>of</strong> the field at two adjacent elements. . . . . . . 90<br />

4.2 Transformation <strong>of</strong> the physical quadrilateral to the standard quadrilateral<br />

element configuration (η 1 , η 2 ∈ [−1, 1]). . . . . . . . . . . . . . 96<br />

4.3 Transformation <strong>of</strong> the physical domain triangle to the standard triangular<br />

element (ξ 1 , ξ 2 ∈ [−1, 1] with ξ 1 + ξ 2 ≤ 1) and standard<br />

quadrilateral element configuration (η 1 , η 2 ∈ [−1, 1]). . . . . . . . . . 97<br />

4.4 Basis functions over the standard square element: (a) (p, q) = (0, 0),<br />

(b) (p, q) = (1, 0), (c) (p, q) = (2, 0), (d) (p, q) = (0, 1), (e) (p, q) =<br />

(1, 1), (f) (p, q) = (2, 1), (g) (p, q) = (0, 2), (h) (p, q) = (1, 2), (i)<br />

(p, q) = (2, 2). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98<br />

4.5 Basis functions over the standard triangular region: (a) (p, q) = (0, 0),<br />

(b) (p, q) = (1, 0), (c) (p, q) = (2, 0), (d) (p, q) = (0, 1), (e) (p, q) =<br />

(1, 1), (f) (p, q) = (0, 2). . . . . . . . . . . . . . . . . . . . . . . . . . . 99<br />

6.1 Mixed-element mesh patch for estimating parameter M. . . . . . . . 130


LIST OF FIGURES 17<br />

7.1 Computational mesh (blue lines) and visualization mesh (black lines)<br />

for a P 4 expansion <strong>of</strong> the solution over triangular and quadrilateral<br />

elements. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141<br />

7.2 Convergence rate for the DG discretization <strong>of</strong> the Euler equations. . . 142<br />

7.3 Convergence rate for the DG discretization <strong>of</strong> the Navier-Stokes equations<br />

with the LDG method. . . . . . . . . . . . . . . . . . . . . . . . 144<br />

7.4 Mach contours for the inviscid flow over a cylinder using discontinuous<br />

output <strong>of</strong> the solution and the curvature based boundary conditions<br />

in [92]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145<br />

7.5 Comparison between the numerically evaluated pressure coefficient<br />

and the analytical solution for the inviscid flow around a cylinder. . . 146<br />

7.6 Convergence history for the residual in the density field for the inviscid<br />

flow over a cylinder. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147<br />

7.7 Meshes for the two dimensional Sod’s shock tube problem. . . . . . . 148<br />

7.8 Limited elements in time for the Sod’s shock tube problem using the<br />

quadrilateral mesh <strong>of</strong> Fig. 7.7a. . . . . . . . . . . . . . . . . . . . . . 148<br />

7.9 Comparison <strong>of</strong> the exact density and entropy variation at t = 0.25<br />

with the two-dimensional numerical solution for the Sod’s shock tube<br />

problem in the triangular channel mesh <strong>of</strong> Fig. 7.7b. . . . . . . . . . 149<br />

7.10 Density plot for the Sod’s problem with a pressure ratio <strong>of</strong> 100000. . 150<br />

7.11 Pressure plot for the Sod’s problem with a pressure ratio <strong>of</strong> 100000. . 150<br />

7.12 Entropy plot for the Sod’s problem with a pressure ratio <strong>of</strong> 100000. . 150<br />

7.13 Quadrilateral mesh for supersonic flow at Mach 2 around a cylinder. . 151<br />

7.14 Triangular mesh for supersonic flow at Mach 2 around a cylinder. . . 152<br />

7.15 Mixed type mesh for supersonic flow at Mach 2 around a cylinder. . . 152<br />

7.16 Pressure contour lines using 30 equally spaced intervals for flow at<br />

Mach 2 over a cylinder for the numerical solution obtained with a P 3<br />

approximation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153<br />

7.17 Limited elements for the flow at Mach 2 over a cylinder for the numerical<br />

solution obtained with a quadrilateral mesh and a P 3 approximation;<br />

the blue elements are limited the adjacent green elements are<br />

P 1 expansions, the adjacent to green red elements are P 2 , and for the<br />

rest <strong>of</strong> the domain P 3 expansions are employed. . . . . . . . . . . . . 154


LIST OF FIGURES 18<br />

7.18 Limited elements for the flow at Mach 2 over a cylinder for the numerical<br />

solution obtained with a P 2 approximation; the blue elements<br />

are limited the adjacent red elements are P 1 expansions and for the<br />

rest <strong>of</strong> the domain P 2 expansions are employed. . . . . . . . . . . . . 155<br />

7.19 Comparison <strong>of</strong> the pressure distribution along the stagnation line obtained<br />

from P 2 and P 3 approximations <strong>of</strong> the numerical solutions<br />

with different meshes. . . . . . . . . . . . . . . . . . . . . . . . . . . . 156<br />

7.20 Variation <strong>of</strong> the maximum value <strong>of</strong> parameter M for each characteristic<br />

field obtained during convergence to steady state <strong>of</strong> the numerical<br />

solution for the flow at Mach 2 over a cylinder with a quadrilateral<br />

mesh and P 2 solution expansion. . . . . . . . . . . . . . . . . . . . . 157<br />

7.21 Limited elements on the coarse quadrilateral mesh for the double<br />

mach reflection problem at t = 0.2. . . . . . . . . . . . . . . . . . . . 158<br />

7.22 Density contours on the coarse quadrilateral mesh for the double mach<br />

reflection problem at t = 0.2. . . . . . . . . . . . . . . . . . . . . . . . 159<br />

7.23 Density contours for the numerical solution computed on the rectangular<br />

mesh, h = 1/240, and a P 1 approximation for the double mach<br />

reflection problem. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159<br />

7.24 Limited elements for the numerical solution computed on the rectangular<br />

mesh, h = 1/240, and a P 1 approximation for the double<br />

mach reflection. The blue elements are limited and for the rest <strong>of</strong> the<br />

domain a P 1 approximation <strong>of</strong> the solution is employed. . . . . . . . 160<br />

7.25 Limited elements for the numerical solution computed on the triangular<br />

mesh, h = 1/240, and a P 1 approximation for the double mach<br />

reflection. The blue elements are limited and for the rest <strong>of</strong> the domain<br />

a P 1 approximation <strong>of</strong> the solution is employed. . . . . . . . . . 161<br />

7.26 Density contours for the numerical solution computed on the triangular<br />

mesh, h = 1/240, and a P 1 approximation for the double mach<br />

reflection problem. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161<br />

7.27 Density contours for the numerical solution computed on the triangular<br />

mesh, h = 1/240, and a P 2 approximation for the double mach<br />

reflection problem. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162<br />

7.28 Variation <strong>of</strong> the maximum value <strong>of</strong> parameter M for each characteristic<br />

field obtained during the time advancement <strong>of</strong> the numerical<br />

solution for the double mach reflection problem obtained with a uniform<br />

triangular element mesh with a P 2 approximation and h = 1/240.163


LIST OF FIGURES 19<br />

7.29 Limited elements for the numerical solution obtained on a rectangular<br />

mesh and P 1 approximation and h = 1/80 for the flow at Mach<br />

number <strong>of</strong> 3 in a wind tunnel with a forward facing step at t = 2.0. . 164<br />

7.30 Density field for the numerical solution obtained on a rectangular<br />

mesh and P 1 approximation and h = 1/80 for the flow at Mach<br />

number <strong>of</strong> 3 in a wind tunnel with a forward facing step at t = 2.0. . 165<br />

7.31 Limited elements for the numerical solution obtained on a fine triangular<br />

mesh and a P 2 approximation and h = 1/100 for the flow<br />

at M = 3.0 in a wind tunnel with a forward facing step at t = 4.0.<br />

Blue elements: limited solution. Green elements P 1 expansion. Red<br />

elements P 2 expansion. . . . . . . . . . . . . . . . . . . . . . . . . . . 165<br />

7.32 Density field for the numerical solution obtained on a fine triangular<br />

mesh and P 2 approximation and h = 1/100 for the flow at Mach<br />

number <strong>of</strong> 3 in a wind tunnel with a forward facing step at t = 4.0. . 166<br />

7.33 Mixed-type element mesh for the inviscid flow at Mach number <strong>of</strong> 3<br />

in a tunnel with a step. . . . . . . . . . . . . . . . . . . . . . . . . . . 166<br />

7.34 Density field for the numerical solution obtained on a mixed-type<br />

element mesh and a P 1 approximation for the flow at Mach number<br />

<strong>of</strong> 3 in a wind tunnel with a forward facing step at t = 4.0. . . . . . . 167<br />

7.35 Limited elements at t = 0.1 using a mixed-type element mesh for the<br />

inviscid flow at Mach number <strong>of</strong> 3 in a wind tunnel with a forward<br />

facing step. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167<br />

7.36 Variation <strong>of</strong> the parameter M for the TVB limiter applied on the<br />

mixed-type element mesh for the inviscid flow at Mach number <strong>of</strong> 3<br />

in a wind tunnel with a forward facing step. . . . . . . . . . . . . . . 168<br />

7.37 Sample meshes with density contours at t = 0.8 for the diffraction<br />

<strong>of</strong> a strong shock over a backward facing step using a P 2 solution<br />

expansion. The black square shows where the computational domain<br />

was clipped for showing the underlying computational mesh. . . . . . 169<br />

7.38 Limited elements for the diffraction <strong>of</strong> strong shock over a backward<br />

facing step at t = 2.0. Blue elements: limited solution. Green elements<br />

P 1 expansion. Red elements P 2 expansion. . . . . . . . . . . . 170<br />

7.39 Density contours for the diffraction <strong>of</strong> strong shock over a backward<br />

facing step at t = 2.0 with 30 equally spaced intervals from 0.06 to<br />

7.16. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171


LIST OF FIGURES 20<br />

7.40 Pressure contours for the diffraction <strong>of</strong> strong shock over a backward<br />

facing step at t = 2.0 with 30 equally spaced intervals from 0.09 to<br />

31.9. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172<br />

7.41 Plot <strong>of</strong> density and pressure along the line defined by points [0, 7] and<br />

[13, 7]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173<br />

7.42 Numerically computed pressure (left) and density (right) gradient<br />

field for the diffraction <strong>of</strong> strong shock over a backward facing step<br />

at t = 2.0, using a rectangular elements mesh. . . . . . . . . . . . . . 173<br />

7.43 Numerically computed pressure (left) and density (right) gradient<br />

field for the diffraction <strong>of</strong> strong shock over a backward facing step<br />

at t = 2.0, using a triangular elements mesh. . . . . . . . . . . . . . . 174<br />

7.44 Variation <strong>of</strong> the second derivative estimate for the diffraction <strong>of</strong> a<br />

strong shock over a backward facing step using a P 2 expansion on a<br />

rectangular elements mesh. . . . . . . . . . . . . . . . . . . . . . . . . 174<br />

7.45 Plot <strong>of</strong> density and pressure along a line defined by the points [2, 5.5]<br />

and [5, 4] at time t = 0.9. . . . . . . . . . . . . . . . . . . . . . . . . . 175<br />

7.46 Vorticity field for the diffraction <strong>of</strong> a Mach 5 shock over a backward<br />

facing step. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177<br />

7.47 Density contours for Schardin’s problem with 30 equal spaced intervals<br />

from 0.6 to 2.4 using a P 2 solution expansion . . . . . . . . . . . 179<br />

7.48 Pressure contours for Schardin’s problem with 30 equal spaced intervals<br />

from 0.4 to 2.14 using a P 2 expansion. . . . . . . . . . . . . . . . 180<br />

7.49 Numerical Schlieren for Schardin’s problem using a P 2 solution expansion.<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181<br />

7.50 Shadowgraph for the Schardin’s problem at t = 128µs. . . . . . . . . 182<br />

7.51 Numerical Schlieren for the Shardin’s problem on the finest mesh and<br />

using P 2 solution expansion. . . . . . . . . . . . . . . . . . . . . . . . 183<br />

7.52 Numerical Schlieren for Schardin’s problem on the refined triangular<br />

mesh with mesh element size in the vortex region equal to h 1 = 0.001. 184<br />

7.53 Numerical Schlieren for Schardin’s problem on the refined triangular<br />

mesh with mesh element size in the vortex region equal to h 2 = 0.0005.185<br />

7.54 Density contours using a P 2 expansion for the flow <strong>of</strong> a strong shock<br />

moving at Mach number <strong>of</strong> 5 around a triangle with 30 equal spaced<br />

intervals from 0.28 to 13.28. Results are depicted at non-dimensional<br />

time. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187


LIST OF FIGURES 21<br />

7.55 Limited elements for the flow <strong>of</strong> a strong shock moving at Mach number<br />

<strong>of</strong> 5 around a triangle along with the gradient <strong>of</strong> pressure over 30<br />

equally spaced intervals from 0 to 1000. Results are depicted at nondimensional<br />

time. Blue elements: limited solution. Green elements<br />

P 1 expansion. Red elements P 2 expansion. . . . . . . . . . . . . . . . 188<br />

7.56 Comparison <strong>of</strong> the U-velocity pr<strong>of</strong>ile with the Blasius solution for the<br />

flow over a flat plate. . . . . . . . . . . . . . . . . . . . . . . . . . . . 190<br />

7.57 Comparison <strong>of</strong> the V-velocity pr<strong>of</strong>ile with the Blasius solution for the<br />

flow over a flat plate. . . . . . . . . . . . . . . . . . . . . . . . . . . . 190<br />

7.58 Velocity field vectors and vorticity field for the flat plate boundary<br />

layer solution using a P 3 solution expansion. . . . . . . . . . . . . . . 191<br />

7.59 Visualization mesh for a P 4 computation for the two tandem NACA<br />

0012 airfoils. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192<br />

7.60 Lift and drag coefficients for the forward airfoil. . . . . . . . . . . . . 194<br />

7.61 Lift and drag coefficients for the aft airfoil. . . . . . . . . . . . . . . . 195<br />

7.62 Velocity contours for the flow over NACA0012 tandem airfoils. . . . . 196<br />

7.63 Vorticity contours for the flow over NACA0012 tandem airfoils. . . . 196


List <strong>of</strong> Tables<br />

2.1 Reference quantities used for the non-dimensionalization <strong>of</strong> the Navier-<br />

Stokes equations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46<br />

3.1 Expressions for the error coefficients for each type <strong>of</strong> mesh distortion. 66<br />

3.2 Expressions for the error coefficients for the mixed-type element mesh<br />

interfaces. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67<br />

3.3 Central interface point displacements for improving the accuracy <strong>of</strong><br />

the u y derivative evaluation for the mixed-type element mesh interfaces. 67<br />

3.4 Characteristic local lengths for the mixed-type element mesh interfaces. 70<br />

3.5 Mesh quality index Q expressions for the elementary types <strong>of</strong> mesh<br />

distortion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72<br />

4.1 Basis functions for a third-order approximation over the standard<br />

quadrilateral region. . . . . . . . . . . . . . . . . . . . . . . . . . . . 99<br />

4.2 Basis functions for a third-order approximation over the standard<br />

triangular region. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99<br />

4.3 Types <strong>of</strong> Gauss numerical integration. . . . . . . . . . . . . . . . . . . 102<br />

4.4 DG methods and their interiors fluxes taken for Arnold et. al [6]. . . 108<br />

5.1 Butcher’s table . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119<br />

7.1 Results for the sensitivity study regarding the far field boundary distance.<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193<br />

22


Chapter 1<br />

Introduction<br />

The present work focuses on the numerical solution <strong>of</strong> the compressible flow equations<br />

for flows with strong moving shocks and in the a priori evaluation <strong>of</strong> the<br />

geometrical quality <strong>of</strong> the mesh used for the discretization <strong>of</strong> the computational domain.<br />

The high-order Discontinuous Galerkin (DG) method is used for discretizing<br />

the system <strong>of</strong> the compressible Navier-Stokes equations in space along with highorder<br />

explicit and implicit discretization schemes for unsteady problems, in order to<br />

achieve high spatial and temporal resolution <strong>of</strong> the complex flow features appearing<br />

in flows with strong moving shocks.<br />

The numerical solution <strong>of</strong> the equations <strong>of</strong> compressible flow is an extremely<br />

powerful technique for the analysis <strong>of</strong> flows over a wide range <strong>of</strong> speeds around complex<br />

geometries. Until the mid 1980s, two methodologies prevailed the study <strong>of</strong> fluid<br />

motion, namely the experimental and the theoretical methodology. The former uses<br />

and develops techniques for the understating <strong>of</strong> the physical phenomena associated<br />

with a flow field. However, experiments are costly and time consuming and in some<br />

cases, even impossible to perform due to technical difficulties and physical limitations.<br />

An example <strong>of</strong> such cases is the study <strong>of</strong> the flow field over a re-entry vehicle.<br />

The theoretical approach attempts to solve the governing equations <strong>of</strong> fluid flow, or<br />

simplified versions <strong>of</strong> them, using mathematical methods and advanced numerical<br />

algorithms since analytical solutions exist only for very few simple cases <strong>of</strong> limited<br />

23


CHAPTER 1. INTRODUCTION 24<br />

practical interest.<br />

The advent <strong>of</strong> the digital computer and its enormous development over the<br />

last 50 years both in terms <strong>of</strong> speed and memory capabilities, lead to the creation<br />

<strong>of</strong> a third methodology as a branch <strong>of</strong> the theoretical approach for studding fluid<br />

flow, that <strong>of</strong> the numerical simulation, which is known as Computational Fluid Dynamics<br />

(CFD). The numerical solution <strong>of</strong> the governing equations does not have the<br />

limitations <strong>of</strong> the experimental and theoretical methodology. However, it innately<br />

possesses problems originating from the philosophy and practice <strong>of</strong> the numerical<br />

solution.<br />

The philosophy behind a CFD solution to a specific problem is to ”break”<br />

its mathematical continuum representation into a discrete form and numerically<br />

compute a solution. However, this approach possesses an error and the factors that<br />

contribute to this are the following:<br />

ˆ Discretization <strong>of</strong> the system <strong>of</strong> equations <strong>of</strong> fluid flow.<br />

ˆ Discrete representation <strong>of</strong> the geometry <strong>of</strong> the body around which the flow<br />

field is simulated.<br />

ˆ Round-<strong>of</strong>f errors in computer arithmetic and floating point representation <strong>of</strong><br />

numbers.<br />

ˆ Initial conditions and boundary conditions.<br />

ˆ Iterative solution procedures, which are stopped after some specified tolerances<br />

have been met.<br />

ˆ Modeling errors.<br />

Iterative solution procedures are applied when an implicit time discretization<br />

<strong>of</strong> the nonlinear system describing the motion <strong>of</strong> fluids is performed, but they are<br />

terminated after predefined tolerances are met leading to solutions which are close to<br />

the true solution <strong>of</strong> the system. The initial and especially the boundary conditions<br />

have a pr<strong>of</strong>ound impact on the numerical results and their choice and implementation


CHAPTER 1. INTRODUCTION 25<br />

may be the key for the success <strong>of</strong> a numerical solution for a given flow problem. The<br />

limitations imposed by the computer technology in arithmetic operations, requires<br />

special attention on how the mathematical operations are performed in a computer<br />

code. The discrete representation <strong>of</strong> the geometry in a computer model means that<br />

practically one is solving a problem over a polyhedral and not on a curved continuum<br />

geometry as it is in reality. Simplifications made for the mathematical form <strong>of</strong> the<br />

governing equations in order to be practical to solve them for a setting, simply means<br />

that an accurate physical model is not used. Taking all these factors along with the<br />

discretization methodology employed for ”breaking” the continuum approach, it is<br />

evident that their impact on the discretization process has to be accounted for in<br />

order to a priori minimize, as much as possible, their effect on the final solution.<br />

Emphasis on achieving predictive simulations as demonstrated in [94, 109] has<br />

caused research in numerical simulation to focus on the solution algorithms and the<br />

discretization error. The present research work aims at providing tools that increase<br />

the fidelity <strong>of</strong> the discretization <strong>of</strong> the equations <strong>of</strong> gas dynamics. The Finite Volume<br />

(FV) method is currently wide spread in CFD, however its successful application<br />

strongly depends on the quality <strong>of</strong> the mesh used for discretizing the physical space.<br />

For those reasons, a mesh induced error analysis based on a novel approach <strong>of</strong><br />

the present thesis has been conducted in order to equip the CFD community with<br />

better and more accurate metrics for the appropriateness <strong>of</strong> a mesh for a CFD<br />

solution using the FV method. The present work assesses grid quality by computing<br />

appropriate metrics <strong>of</strong> the elements. However, there are two distinguishing aspects<br />

<strong>of</strong> it: (i) the metrics indicating quality are derived directly from related analytic<br />

forms <strong>of</strong> the truncation error and (ii) can give analytic expressions for reducing<br />

the discretization error via re-shaping <strong>of</strong> the elements. The primary issue with TE<br />

analysis is the complexity <strong>of</strong> the related expressions, especially for multi-dimensions<br />

and for general mixed-type element mesh topologies. The present work addresses<br />

this complexity barrier via employment <strong>of</strong> symbolic mathematics s<strong>of</strong>tware [101].<br />

The present work analyzes a generally-distorted mesh (structured, unstructured<br />

and mixed-type) into elementary distortions (stretching, skewness, shearing,<br />

rotation), as well as three common types <strong>of</strong> interfaces, and those were directly related


CHAPTER 1. INTRODUCTION 26<br />

to the TE. The derived analytic expressions are relatively simple and amenable to<br />

future work on improving the mesh. The derived quality indices can be also applied<br />

for assessing the mesh quality for a higher order discretization method as demonstrated<br />

in [99].<br />

Moreover, due to the limits <strong>of</strong> the FV method in terms <strong>of</strong> resolution, especially<br />

for flows with complex features with strong moving shocks and the difficulty in generating<br />

a good structured grid around complex geometries for Finite Difference (FD)<br />

discretizations, a DG code has been developed and named HoAc, which stands<br />

for High Order Accuracy. The developed code uses PETSc [58, 59] for performing<br />

parallel computations using the domain decomposition method. It can handle<br />

mixed-type element meshes employing hierarchical modal basis functions [83] up to<br />

arbitrary order <strong>of</strong> accuracy, for the discretization <strong>of</strong> the compressible Navier-Stokes<br />

equations. Explicit and implicit time schemes are used provided by PETSc and a<br />

novel shock capturing procedure has been developed, implemented and thoroughly<br />

tested. All the computations are performed in the standard element configuration,<br />

thus eliminating the complexity and difficulties arising by direct application <strong>of</strong> the<br />

solution algorithms in the physical space.<br />

The novel shock capturing approach proposed in this work uses the basic TVB<br />

limiter proposed by Cockburn and Shu [82]. However, it overcomes the ambiguity<br />

associated with the a priori specification <strong>of</strong> a parameter that estimates the second<br />

derivative <strong>of</strong> the solution. Application <strong>of</strong> the TVB limiter for the canonical elements<br />

<strong>of</strong> the computational domain, in a dimension per dimension fashion, renders our<br />

approach essentially the same with the original TVB limiter [82] that until now<br />

was used for rectangular elements. For the canonical element <strong>of</strong> the computational<br />

domain, tensor product orthogonal polynomial bases are constructed as described in<br />

detail in [83]. However, in contrast to the original TVB limiter, the second derivative<br />

<strong>of</strong> the numerical solution is estimated for each field, and it is not set to a constant<br />

value, (as suggested in the original paper [82]). Note that in the original approach<br />

[82] <strong>of</strong>ten and for cases with strong shocks and large pressure ratios, the suggested<br />

value must be altered, while in our approach no user intervention is required. Recent<br />

applications for flows with strong discontinuities resulting from high pressure ratios


CHAPTER 1. INTRODUCTION 27<br />

[156] have demonstrated that the constant value <strong>of</strong> the second derivative, needs to<br />

be arbitrarily readjusted for each field in order to avoid divergence. Application <strong>of</strong><br />

the limiter in the transformed domain has the additional advantage that limiting for<br />

rectangular, quadrilateral, and triangular elements is essentially the same, since the<br />

inverse transformation and the collapsed coordinates system introduced in [83] is<br />

used to recover the element in the physical space. The proposed limiting approach<br />

is accurate, effective, and can be applied on distorted mixed-type meshes without<br />

readjusting parameters for a wide range <strong>of</strong> flow problems discussed in the results<br />

section. The extension <strong>of</strong> the proposed limiting approach for three dimensions is<br />

straightforward [90].<br />

In the following sections <strong>of</strong> the introduction chapter, the classical methods used<br />

in CFD are described along with comments on what extent they are affected by the<br />

underlying mesh employed for their application, and their ability to extend them in<br />

orders <strong>of</strong> accuracy higher than two. Furthermore, the selection <strong>of</strong> the discretization<br />

method employed in the current work is advocated and the overview <strong>of</strong> the thesis is<br />

presented.<br />

1.1 Classical methods in CFD<br />

Discretization techniques have been developed since the early 1900s due to the numerous<br />

efforts for developing algorithms for the numerical solution <strong>of</strong> partial differential<br />

equations (PDEs). Currently, the most popular and well established methods<br />

in CFD are the Finite Difference (FD) [97], the Finite Volume (FV) [96] and the<br />

Finite Element (FE) methods [83], and most CFD methods used in practice are at<br />

most second-order accurate in space [145, 148]. A brief description <strong>of</strong> each method<br />

follows.<br />

Finite Difference methods<br />

The most basic discretization technique for a system <strong>of</strong> PDEs is that <strong>of</strong> the FD<br />

method. A structured canonical grid, through the use <strong>of</strong> a generalized coordinate


CHAPTER 1. INTRODUCTION 28<br />

transformation is employed, where the method approximates the analytical form <strong>of</strong><br />

the derivatives appearing in the differential equations with finite differences leading<br />

to a discrete representation <strong>of</strong> the differential form, which can be solved numerically.<br />

There exists a plethora <strong>of</strong> approaches for the discrete estimation <strong>of</strong> the derivatives<br />

and some excellent text books describing these methodologies are those from Hirsch<br />

[63, 64], Anderson [4] and Tannehill et al. [132] to mention a few.<br />

The main advantages <strong>of</strong> the FD methods are that they are easy to program<br />

and efficient in terms <strong>of</strong> computational cost. Due to their efficiency and simplicity<br />

they have been used for computationally intensive problems. It is possible to<br />

construct high-order extensions <strong>of</strong> the FD methods, by adding more grid points in<br />

the computational stencil, since the accuracy <strong>of</strong> the method is determined by the<br />

estimation <strong>of</strong> the derivatives [54, 61, 95, 147]. However, in order to retain high-order<br />

accurate numerical solutions with FD discretizations it is required to have smooth<br />

canonical meshes in the physical domain.<br />

A major drawback, however, <strong>of</strong> the FD method is that it strictly requires structured<br />

grids for its application, thus limiting its use on simple geometrical configurations.<br />

It is possible to apply the method for problems accompanied with complex<br />

geometries. However, this is a difficult task, which practically turns the FD method<br />

inappropriate for computing flows around complex configurations. In principle it is<br />

possible to construct a FD scheme on unstructured grids [98], but this requires a<br />

reconstruction <strong>of</strong> a polynomial function from grid points, which is a complex problem<br />

even in the two dimensional case. Moreover, the high-order versions <strong>of</strong> the FD<br />

method require a regular, smooth grid for stability and accuracy reasons. Consequently,<br />

FD methods are mostly used for fundamental studies on simple geometries.<br />

Usually, in order to ensure stability <strong>of</strong> the scheme the order is reduced near and at<br />

the boundaries, which results in a special grid refinement close to the boundaries so<br />

that to compensate for the loss <strong>of</strong> accuracy, or use <strong>of</strong> one sided differencing schemes<br />

that must be employed in these regions.


CHAPTER 1. INTRODUCTION 29<br />

Finite Volume methods<br />

The FV method is quite popular in the CFD community. The starting point for a<br />

FV discretization <strong>of</strong> a system <strong>of</strong> partial differential equations is the integral form<br />

<strong>of</strong> the system. A tessellation <strong>of</strong> the physical space in elements is performed and in<br />

each element the governing equations are discretized. In contrast to the FD method,<br />

where derivatives are being estimated, fluxes through the element boundaries are<br />

evaluated. There is a plethora <strong>of</strong> choices <strong>of</strong> how these fluxes can be chosen. A very<br />

popular approach for hyperbolic type systems is the upwind method [64, 96] where<br />

the flux estimate is based on the structure <strong>of</strong> the local wave propagation, leading<br />

to robust solution algorithms. Another approach, which however requires an ad hoc<br />

specification <strong>of</strong> parameters for stability reasons and shock capturing, is the artificial<br />

dissipation method [69].<br />

One significant advantage <strong>of</strong> the FV method is its ability to perform well on<br />

structured, unstructured, and mixed-type element meshes in two and three dimensions.<br />

This permits the application <strong>of</strong> the method on problems involving complex<br />

geometries. Formally, the method is first order accurate, but with a reconstruction<br />

procedure [14, 96] <strong>of</strong> the solution at the element boundaries, its accuracy may be<br />

extended to higher order. However, in practice this reconstruction is quite computationally<br />

expensive and difficult to perform, even for moderate orders <strong>of</strong> accuracy<br />

(third-order for instance) [47], and most FV codes on unstructured meshes have<br />

resolution properties belonging to the second-order accurate schemes.<br />

The FV as the FD method requires the construction <strong>of</strong> regular and smooth<br />

meshes. Moreover, due to the difficulty in constructing a higher order FV scheme,<br />

use <strong>of</strong> the second-order accurate FV method for practical flow problems necessitates<br />

the construction <strong>of</strong> fine meshes especially for the solution <strong>of</strong> the Navier-Stokes<br />

equations.


CHAPTER 1. INTRODUCTION 30<br />

Finite Element methods<br />

FE methods use a tessellation <strong>of</strong> the physical space in elements where the weak<br />

form <strong>of</strong> the governing equations is solved. Specifically, the weak form is derived<br />

by multiplying the system <strong>of</strong> the PDEs with a weighting function and integrating<br />

over the elemental space. Then, a polynomial expansion <strong>of</strong> the solution is chosen<br />

and the discrete from <strong>of</strong> the system is derived. The choice <strong>of</strong> the test (also known<br />

as weighting) and the expansion functions dictates which type <strong>of</strong> FE method is<br />

being employed. Versions <strong>of</strong> FE methods are the Galerkin [21, 24, 83], also known<br />

as the Bubnov-Galerkin method, the Petrov-Galerkin [62, 65], also known as the<br />

generalized Galerkin method and the least-squares FE methods [114, 115]. Highorder<br />

versions for FE methods can be constructed by increasing the order <strong>of</strong> the<br />

polynomial used for the solution expansion. Increase <strong>of</strong> the order <strong>of</strong> accuracy in<br />

FE methods comes however at the expense <strong>of</strong> computing cost, algorithmic and<br />

programming complexity.<br />

The Bubnov-Galerkin method uses the same function space both for the test<br />

function and the solution expansion, leading to a symmetric representation <strong>of</strong> the<br />

weak formulation, and it is applicable to any type <strong>of</strong> partial differential equations.<br />

The Petrov-Galerkin version <strong>of</strong> the FE method is used for solving partial differential<br />

equations with odd terms. In this version the formulation method is the<br />

same as the normal Galerkin method. However, the test function and the solution<br />

function approximations are chosen not to be the same, hence they must be approximated<br />

separately and thus, the representation <strong>of</strong> the weak form looses symmetry.<br />

The least squares FE method originates from the idea <strong>of</strong> least squares approximation<br />

developed by Gauss. In this version <strong>of</strong> the FE method the residual <strong>of</strong> the<br />

system obtained from the solution expansion is set as the test function, leading to<br />

a symmetric form <strong>of</strong> the weak representation.<br />

FE methods may be classified into two categories: continuous and discontinuous.<br />

Continuous FE methods employ an expansion set, which is continuous over<br />

the entire domain <strong>of</strong> the physical space being tessellated into non overlapping elements.<br />

On the other hand, discontinuous FE methods perform an approximation <strong>of</strong>


CHAPTER 1. INTRODUCTION 31<br />

the solution <strong>of</strong> the system <strong>of</strong> PDEs at each element separately permitting it to be<br />

discontinuous across inter element boundaries.<br />

Higher-order versions <strong>of</strong> the FE method do not exhibit strong dependence on<br />

the mesh constructed for a field simulation. This is due to the local support <strong>of</strong> the<br />

solution from the local expansion at every element. In the present work, the DG<br />

method which belongs to the version <strong>of</strong> Bubnov-Galerkin FE methods is used and<br />

it will be presented in detail in a following chapter.<br />

1.2 Why the DG method<br />

Accurate prediction <strong>of</strong> aerodynamic and thermal loads on aerospace vehicles is a crucial<br />

stage <strong>of</strong> the design cycle. Wind tunnel investigations and theoretical predictions<br />

are used to estimate these loads. Experimental measurements <strong>of</strong> high speed flows<br />

are costly, difficult to perform, and usually unable to reach realistic flight Reynolds<br />

numbers. On the other hand, highly accurate theoretical predictions with DNS or<br />

LES are limited to simple flow configurations [2] and low Reynolds numbers. Most<br />

<strong>of</strong> these calculations are performed with structured grids that are more efficient<br />

computationally for high resolution simulations. For flows without discontinuities,<br />

high-order compact FD schemes [54, 147] can be used to provide the necessary highorder<br />

spatial accuracy at a reasonable computing cost. On the other hand, flows with<br />

strong discontinuities can be efficiently computed on structured grids using compact<br />

schemes with filters [131, 153] or the FD ENO [73, 130] and weighted ENO (WENO)<br />

[128] discretizations. The extension <strong>of</strong> ENO [1] or WENO [110] methodology to unstructured<br />

meshes is, however, not straightforward and becomes computationally<br />

intensive. Similarly, the high-order FV methods for three dimensional calculations<br />

are quite intensive in memory and computational time, and difficult to implement<br />

[47].<br />

Use <strong>of</strong> standard second-order accurate in space FV and FD methods for the<br />

prediction <strong>of</strong> aerodynamic and thermal loads over more complex configurations, as<br />

mentioned before, is currently widespread. These methods, however, require very


CHAPTER 1. INTRODUCTION 32<br />

high grid densities for the accurate computation <strong>of</strong> the complex flow features such<br />

as those generated from the interaction <strong>of</strong> shock waves with boundary layers even<br />

on relatively simple geometrical configurations. Use <strong>of</strong> unstructured meshes could<br />

reduce computing requirements for accurate computation <strong>of</strong> complex shock interactions,<br />

because they allow selective refinement in critical flow regions. Recently, it<br />

was demonstrated [149], that for these flows only the combination <strong>of</strong> selective grid<br />

refinement and high-order accuracy can achieve the desired level <strong>of</strong> accuracy at a<br />

reduced computing cost for complex shock wave interactions. In addition, Shi et<br />

al. [128] clearly demonstrated that resolution <strong>of</strong> complex flow features, generated<br />

by the impingement <strong>of</strong> a strong shock on a compression ramp, could be equally well<br />

achieved by grid refinement (doubling the resolution along each coordinate direction)<br />

or with the increase <strong>of</strong> the order <strong>of</strong> accuracy. The results <strong>of</strong> Shi et al. [128],<br />

obtained with the WENO scheme, indicate that increase <strong>of</strong> the order <strong>of</strong> accuracy<br />

<strong>of</strong> the spatial discretization scheme by an order <strong>of</strong> two is equivalent to doubling the<br />

resolution <strong>of</strong> the mesh in every direction, as far as the resolution <strong>of</strong> smooth but<br />

complex flow features is concerned. Similar conclusions concerning the enhancement<br />

<strong>of</strong> resolution were drawn for high-order discretizations with the DG method<br />

[53, 137–139] and the closely related Spectral Volume (SV) method [5, 149].<br />

The main advantages <strong>of</strong> the DG [28, 41, 44] method compared to other available<br />

high-order methods is that, as FE type method, it has a compact stencil, it is<br />

applicable to arbitrary unstructured meshes, and it can handle hanging nodes and<br />

curved geometries in a natural manner. Therefore, this method is suitable for grid<br />

adaptation, and appropriate for the computation <strong>of</strong> flows with strong discontinuities.<br />

These significant advantages <strong>of</strong> the DG method over other high-order methods<br />

come unfortunately at the expense <strong>of</strong> implementation complexity, higher computing<br />

cost, and reduction <strong>of</strong> the allowable, by stability limitations, time step for marching<br />

in time. Computation in regions <strong>of</strong> strong shocks and other discontinuities with<br />

the DG method could be stabilized with the application <strong>of</strong> classical total variation<br />

bounded (TVB) limiters [28, 44]. In regions where the TVB limiter is activated the<br />

formal order <strong>of</strong> accuracy <strong>of</strong> the method drops. More recently, alternative approaches<br />

for the computation <strong>of</strong> flows containing discontinuities with high-order accurate DG


CHAPTER 1. INTRODUCTION 33<br />

space discretizations were proposed by applying filters [100, 138], or by adding numerical<br />

dissipation [105, 142]. However, use <strong>of</strong> these approaches increases the overall<br />

computational cost <strong>of</strong> the DG method and does not guarantee that high accuracy<br />

is retained at the shock, where the computation essentially drops to lower order<br />

<strong>of</strong> accuracy, so that preservation <strong>of</strong> monotonicity is ensured. Filters and artificial<br />

dissipation in addition require ad hoc specification <strong>of</strong> tuning parameters.<br />

High resolution <strong>of</strong> strong shocks can also be achieved with the selective increase<br />

<strong>of</strong> grid resolution (h-refinement) and use <strong>of</strong> lower order spatial discretizations<br />

with TVB limiters at the regions <strong>of</strong> discontinuities. This approach appears quite<br />

appropriate for the resolution <strong>of</strong> shocks with DG discretizations and was applied<br />

successfully for hexahedral meshes in [142]. In addition hp-refinement schemes [103]<br />

have appeared in the literature.<br />

The DG method has become popular in recent years due to its high-order <strong>of</strong><br />

accuracy, and the ability to simulate flows around complex geometries with the ease<br />

<strong>of</strong> increasing the order <strong>of</strong> approximation, while keeping a compact stencil. The DG<br />

method has higher computational cost compared with standard second-order FV<br />

methods. However, the compact stencil <strong>of</strong> the DG method provides an essential advantage<br />

for achieving high-order accuracy and parallelization <strong>of</strong> the algorithms, both<br />

with the use <strong>of</strong> domain decomposition through message passage interface (MPI), and<br />

graphics processor units (GPUs) [50, 81, 87]. In this work, parallelization <strong>of</strong> the DG<br />

method through domain decomposition and use <strong>of</strong> MPI was performed.<br />

Accurate predictions <strong>of</strong> skin friction and thermal loads <strong>of</strong> high speed complex<br />

flows in both simple and nontrivial geometries, require high resolution computations.<br />

High-order DG discretizations possess features that make them very attractive for<br />

computation <strong>of</strong> complex flows with strong shocks. A key ingredient that would<br />

make the DG method more suitable for these computations, is the straightforward<br />

application <strong>of</strong> p-adaptive procedures that ensure robust and accurate capturing <strong>of</strong><br />

discontinuities along with high-order <strong>of</strong> accuracy in the smooth parts <strong>of</strong> the flow. In<br />

this spirit, low order expansions and high mesh density could be used in regions <strong>of</strong><br />

discontinuities. On the other hand, resolution <strong>of</strong> smooth but complex flow features,<br />

such as vortices and shear layers that usually evolve in time, could be obtained


CHAPTER 1. INTRODUCTION 34<br />

with relatively uniform mesh density using higher order expansions. Flow features<br />

where p-adaptivity is required could either be detected in the solution as regions<br />

<strong>of</strong> high gradients, such as density and pressure variations in vortical regions, or by<br />

a-posteriori error estimates <strong>of</strong> the FE solution [10, 108, 144]. However, these aspects<br />

<strong>of</strong> the DG method have not been addressed satisfactorily yet, and utilization <strong>of</strong> p-<br />

adaptivity is not currently widespread. Therefore, the ability <strong>of</strong> the DG method<br />

to compute flows with discontinuities in a unified manner for arbitrary mixed-type<br />

meshes requires further development. Moreover, limiting in three dimensions for<br />

arbitrary unstructured meshes has not been demonstrated. The DG method reduces<br />

the solution accuracy near discontinuities. This affects the resolution properties <strong>of</strong><br />

the method and much <strong>of</strong> its potential could be lost when limiting is extended over<br />

large regions. Therefore, narrowing <strong>of</strong> the regions requiring limiting and use <strong>of</strong> p-<br />

adaptive procedures is expected to enhance the potential and broaden utilization <strong>of</strong><br />

the DG method.<br />

The slope limiters for the computation <strong>of</strong> flows with discontinuities used so<br />

far with the DG method have been demonstrated mainly for rectangular elements.<br />

A total variation bounded (TVB) limiting procedure for the DG method was constructed<br />

for solving scalar, one-dimensional, hyperbolic conservation laws by Shu<br />

[129] and Cockburn and Shu [82]. The extension <strong>of</strong> the method to one-dimensional<br />

systems has been applied with satisfactory results [40]. The basic idea behind the<br />

TVB limiter was to use a slope limiter to preserve the monotonicity <strong>of</strong> the solution<br />

averages. The same nonlinear TVB limiting was also applied for two dimensional<br />

problems [27, 32, 41], for second (P 1 polynomial expansion) and third-order (P 2 )<br />

accurate computations with rectangular meshes. One ambiguity <strong>of</strong> this limiting<br />

procedure is that it relies on a priori estimates <strong>of</strong> the numerical solution’s second<br />

derivative. Moreover, the application for triangular meshes [40, 82] is more elaborate<br />

and straightforward application <strong>of</strong> the method for mixed-type meshes has not<br />

been addressed yet.<br />

In an effort to extend the TVB limiter for higher order accurate solutions, Qiu<br />

and Shu [120] applied the same TVB limiter for the one-dimensional Euler equations,<br />

by reconstructing the solution at every element with a Hermite WENO scheme using


CHAPTER 1. INTRODUCTION 35<br />

the average solution at every element. Applying this idea they achieved a third-order<br />

monotone solution. Similar approaches were followed by Zhu and Qiu [157, 158] and<br />

by Luo et al. [100] for multidimensional calculations. These limiting approaches were<br />

demonstrated for meshes with rectangular elements and for triangular meshes using<br />

the approach suggested by Cockburn [40], which is not straight forward to extend<br />

in three dimensions. In addition, the approach <strong>of</strong> Zhu [157, 158] implies increase <strong>of</strong><br />

the stencil width, and one <strong>of</strong> the main advantages <strong>of</strong> the DG method is lost. Biswas<br />

et al. [29] proposed an alternative implementation <strong>of</strong> limiting, where the solution<br />

moments are altered in order to preserve monotonicity. Krivodonova, [91] applied<br />

this concept with promising results, in computations for flows with strong shocks<br />

and with order <strong>of</strong> accuracy higher than two for rectangular elements. On the other<br />

hand Jaffre [68] and Van der Ven [143] added artificial dissipation to stabilize the<br />

DG method.<br />

From the above it is concluded in the present thesis that the DG method as implemented<br />

and enhanced with the newly developed limiting approach is suitable for<br />

the computation <strong>of</strong> high speed complex flow fields around complex geometries. The<br />

present work has made a significant improvement to the application <strong>of</strong> TVB limiter<br />

leading to a p-adaptive shock capturing scheme with outstanding performance.<br />

1.3 Thesis outline<br />

Chapter 2 includes a brief presentation <strong>of</strong> the equations <strong>of</strong> gas dynamics, along with<br />

some basic thermodynamic variables. The mathematical character <strong>of</strong> the system<br />

<strong>of</strong> Euler and Navier-Stokes equations is briefly presented, which dictates the type<br />

<strong>of</strong> boundary conditions to be used in the far field boundaries <strong>of</strong> the computational<br />

domain and the non-dimensional form <strong>of</strong> the equations, which is used in all the<br />

computations in this work is shown. The mesh induced error study is performed<br />

in Chapter 3 and the derived novel mesh quality measures for mixed-type element<br />

meshes are presented. These measures are subsequently used in the construction<br />

<strong>of</strong> the new limiter. Several applications <strong>of</strong> the quality measures are also presented,<br />

which confirm their ability to inspect the quality <strong>of</strong> a mesh for a FV discretiza-


CHAPTER 1. INTRODUCTION 36<br />

tion. The DG discretization <strong>of</strong> the Euler and Navier-Stokes equations is presented<br />

in Chapter 4, where the expansion bases used in the present work are presented in<br />

detail. Also, the numerical operations are described for performing all the computations<br />

in the computational space along with the transformation <strong>of</strong> the elemental<br />

physical configuration to its standard representation in the computational space.<br />

Furthermore, the Local Discontinuous Galerkin (LDG) methodology is presented<br />

for discretizing the system <strong>of</strong> Navier-Stokes equations in the DG framework. Chapter<br />

5 deals with the time marching <strong>of</strong> the DG discretization <strong>of</strong> the compressible<br />

equations <strong>of</strong> gas dynamics and the limiting procedure developed for shock capturing<br />

is presented in Chapter 6. In Chapter 7 numerical results for problems with strong<br />

moving shocks are presented along with verification cases that confirm the correctness<br />

<strong>of</strong> the implemented DG discretization and the limiting procedure in HoAc.<br />

Finally, in Chapter 8 the conclusions <strong>of</strong> the current thesis are presented.


Chapter 2<br />

Governing equations <strong>of</strong> gas<br />

dynamics<br />

The governing equations <strong>of</strong> inviscid and viscous compressible flow, along with some<br />

basic thermodynamic relations, will be presented. They are expressed in a Eulerian<br />

reference frame and they are based on the following conservation laws:<br />

ˆ Conservation <strong>of</strong> Mass<br />

ˆ Conservation <strong>of</strong> Momentum<br />

ˆ Conservation <strong>of</strong> Energy<br />

and mathematically they are derived by using Reynolds transport theorem. Their<br />

mathematical character will be stated, which defines the specification and application<br />

<strong>of</strong> the boundary conditions to be used for a numerical solution. The presented<br />

equations will be given first in differential form followed by their non-dimensionalization<br />

and concluding with their integral form.<br />

37


CHAPTER 2. GOVERNING EQUATIONS OF GAS DYNAMICS 38<br />

2.1 Euler equations<br />

The motion <strong>of</strong> a compressible fluid in two dimensions, without any viscous and<br />

thermal conduction effects, is described by the system <strong>of</strong> Euler equations. The<br />

differential form <strong>of</strong> this system in Cartesian coordinates is:<br />

L(U) = ∂U<br />

∂t + ∇ · F i(U) = 0, (2.1)<br />

where U is the conservative variable vector or state vector:<br />

⎡ ⎤<br />

ρ<br />

U =<br />

ρu<br />

⎢<br />

⎣ρv<br />

⎥<br />

⎦ , (2.2)<br />

E<br />

and F i (U) = [f i<br />

g i ], is the inviscid flux tensor with vector components:<br />

⎡ ⎤ ⎡ ⎤<br />

ρu<br />

ρv<br />

f i =<br />

ρu 2 + p<br />

⎢<br />

⎣ ρuv<br />

⎥<br />

⎦ , g i =<br />

ρuv<br />

⎢<br />

⎣ ρv 2 + p<br />

⎥<br />

⎦ . (2.3)<br />

(E + p)u<br />

(E + p)v<br />

This set <strong>of</strong> four equations is closed with the equation <strong>of</strong> state for a perfect gas.<br />

2.1.1 Energy<br />

The total energy E per unit volume is the sum <strong>of</strong> the internal and kinetic energy <strong>of</strong><br />

the flow:<br />

E = ρe + 1 2 ρ(u2 + v 2 ). (2.4)


CHAPTER 2. GOVERNING EQUATIONS OF GAS DYNAMICS 39<br />

The term 1 2 ρ(u2 + v 2 ) is the kinetic energy, while e is the internal energy per unit<br />

mass . Internal energy includes translational, rotational and vibrational energy and<br />

other forms <strong>of</strong> energy in more complicated situations. In the present work the gas<br />

is assumed to be in local chemical and thermodynamic equilibrium and the internal<br />

energy is a function <strong>of</strong> pressure and density:<br />

e = e(p, ρ). (2.5)<br />

2.1.2 Polytropic ideal gas<br />

For an ideal gas the internal energy is a function <strong>of</strong> temperature only, e = e(T ).<br />

The temperature T is related to pressure p and density ρ by the ideal gas law:<br />

p = RρT, (2.6)<br />

where R is a constant obtained by dividing the universal gas constant R by the<br />

molecular weight <strong>of</strong> the gas. To a good approximation, the internal energy is simply<br />

proportional to the temperature T :<br />

e = c V T, (2.7)<br />

where c V is the specific heat at constant volume . Gases that obey Eqs. (2.6) and<br />

(2.7) are called calorically or thermally perfect. If energy is added to a fixed volume<br />

<strong>of</strong> a thermally perfect gas, then the change in internal energy and temperature is<br />

related via:<br />

de = c V dT. (2.8)<br />

However, if the gas is allowed to expand at a constant pressure, not all <strong>of</strong> the<br />

energy goes into increasing the internal energy. The work done by expanding the<br />

volume 1 by d( 1) is pd( 1 ) and the following relation is obtained:<br />

ρ ρ ρ


CHAPTER 2. GOVERNING EQUATIONS OF GAS DYNAMICS 40<br />

de + pd( 1 ρ ) = c pdT ⇔ d(e + p ρ ) = c pdT, (2.9)<br />

where c p is the specific heat at constant pressure . The quantity:<br />

h = e + p ρ , (2.10)<br />

is called the (specific) enthalpy <strong>of</strong> the gas . For a polytropic gas, c p is constant, so<br />

Eq. (2.9) yields:<br />

h = c p T, (2.11)<br />

and the enthalpy is simply proportional to the temperature. By the ideal gas law<br />

results:<br />

c p − c V = R. (2.12)<br />

The equation <strong>of</strong> state <strong>of</strong> a polytropic gas turns out to depend only on the ratio <strong>of</strong><br />

the specific heats:<br />

and also called the adiabatic exponent .<br />

γ = c p<br />

c V<br />

, (2.13)<br />

The internal energy in a molecule is typically split up between various degrees<br />

<strong>of</strong> freedom (translational, rotational, vibrational, etc.). How many degrees <strong>of</strong> freedom<br />

exist, depends on the nature <strong>of</strong> the gas. The general principle <strong>of</strong> equipartition<br />

<strong>of</strong> energy requires that the average energy in each <strong>of</strong> these is the same leading to<br />

each degree <strong>of</strong> freedom contributing an average energy <strong>of</strong>:<br />

1<br />

2 kT,


CHAPTER 2. GOVERNING EQUATIONS OF GAS DYNAMICS 41<br />

per molecule, where k is Boltzmann’s constant and is equal to 1.3806503·10 −23 m 2 kg/s 2<br />

in SI units . This gives a total contribution <strong>of</strong>:<br />

a<br />

2 kT,<br />

per molecule if there are a degrees <strong>of</strong> freedom. Multiplying this by n, the number<br />

<strong>of</strong> molecules per unit mass (which depends on the gas), gives:<br />

e = a nkT. (2.14)<br />

2<br />

The product nk is precisely the gas constant R, which in comparison with Eq.<br />

(2.7), gives:<br />

c V = a R. (2.15)<br />

2<br />

From Eq. (2.12), results that:<br />

c p =<br />

(<br />

1 + a )<br />

R, (2.16)<br />

2<br />

which leads to the following expression for the adiabatic constant:<br />

γ = c p<br />

c V<br />

= a + 2<br />

a . (2.17)<br />

For air a = 5 and γ = 7 2<br />

= 1.4. Noting that:<br />

T =<br />

p<br />

Rρ , (2.18)<br />

leads to:<br />

e =<br />

p<br />

(γ − 1)ρ . (2.19)


CHAPTER 2. GOVERNING EQUATIONS OF GAS DYNAMICS 42<br />

Using Eq. (2.19) in Eq. (2.4), gives the common form <strong>of</strong> the total energy for a<br />

thermally perfect gas:<br />

E =<br />

p<br />

γ − 1 + 1 2 ρ(u2 + v 2 ). (2.20)<br />

and the pressure is related to total energy and the conservative variables through<br />

the equation <strong>of</strong> state:<br />

)<br />

p = (γ − 1)<br />

(E − u2 + v 2<br />

. (2.21)<br />

2ρ<br />

2.1.3 Entropy<br />

Entropy is a fundamental thermodynamic quantity and indicates the degree to which<br />

its internal energy is available for producing usable work. The specific entropy is<br />

given by:<br />

( ) p<br />

s = c V log + constant. (2.22)<br />

ρ γ<br />

from which follows :<br />

p = κe ( s<br />

c V ) ρ γ , (2.23)<br />

where κ is a constant.<br />

2.1.4 Mathematical character <strong>of</strong> the Euler equations<br />

The system <strong>of</strong> Euler equations is hyperbolic, because the eigenvalues <strong>of</strong> the directional<br />

Jacobian matrix [64]:<br />

K = ∂F i · k, (2.24)<br />

∂U


CHAPTER 2. GOVERNING EQUATIONS OF GAS DYNAMICS 43<br />

are real and distinct for any direction represented by the unit vector k . The diagonal<br />

eigenvalue matrix is:<br />

⎡<br />

⎤ ⎡ ⎤<br />

uk x + vk y u<br />

Λ =<br />

uk x + vk y<br />

⎢<br />

⎣uk x + vk y + c<br />

⎥<br />

⎦ = u<br />

⎢<br />

⎣u + c<br />

⎥<br />

⎦ , (2.25)<br />

uk x + vk y − c u − c<br />

where u, v are the Cartesian velocity components , u is the contravariant velocity<br />

component along the k direction, and c is the local speed <strong>of</strong> sound given by the<br />

following equation:<br />

c =<br />

√ γp<br />

ρ . (2.26)<br />

The right and left eigenvectors <strong>of</strong> the directional Jacobian matrix in Eq. (2.24)<br />

are the following:<br />

⎡<br />

R =<br />

⎢<br />

⎣<br />

⎤<br />

ρ<br />

ρ<br />

k x 0<br />

2c<br />

2c<br />

ρu<br />

− ρkx<br />

2c 2<br />

ρv<br />

− ρky ⎥<br />

2c 2 ⎦<br />

q 1 k x<br />

uρk<br />

2 y − vρk x R 1 R 2<br />

ρu<br />

uk x ρk y + ρkx<br />

2c 2<br />

ρv<br />

vk x −ρk x + ρky<br />

2c 2<br />

, (2.27)<br />

⎡<br />

L =<br />

⎢<br />

⎣<br />

− ukx<br />

ρ<br />

− ukx<br />

ρ<br />

k x − kx<br />

c 2 γ−1<br />

2 q 2<br />

− ukx<br />

ρ<br />

− vky<br />

ρ<br />

+ vky<br />

ρ<br />

− vky<br />

ρ<br />

+ (γ−1)<br />

2ρc q 2<br />

k xρ<br />

uk x(γ−1)<br />

c 2<br />

vk x(γ−1)<br />

c 2<br />

− kx(γ−1)<br />

c 2<br />

k y<br />

ρ<br />

− kx ρ<br />

0<br />

− u(γ−1)<br />

ρc<br />

k y<br />

ρ<br />

− v(γ−1)<br />

ρc<br />

+ (γ−1) q 2ρc 2 − kx − u(γ−1) − ky − v(γ−1)<br />

ρ ρc ρ ρc<br />

γ−1<br />

ρc<br />

γ−1<br />

ρc<br />

⎤<br />

, (2.28)<br />

⎥<br />

⎦<br />

with :


CHAPTER 2. GOVERNING EQUATIONS OF GAS DYNAMICS 44<br />

q 1 =<br />

√<br />

u2 + v 2<br />

,<br />

2<br />

q 2 = 2(u 2 + v 2 ) − q 1 ,<br />

R 1 = qρ<br />

4c + uρk x<br />

+ vρk y<br />

2 2<br />

R 2 = qρ<br />

4c − uρk x<br />

− vρk y<br />

2 2<br />

2.2 Navier-Stokes equations<br />

+ ρc<br />

2γ − 2 ,<br />

+ ρc<br />

2γ − 2 .<br />

The system <strong>of</strong> Navier-Stokes equations describes real compressible flows, which<br />

posses both heat conduction and viscosity effects. The differential form <strong>of</strong> this<br />

system is:<br />

N (U) = ∂U<br />

∂t + ∇ · F i(U) = ∇ · F v (U, ∇U), (2.29)<br />

where U and F i (U) is the conservative state vector and the inviscid flux tensor<br />

respectively, and F v (U) = [f v<br />

vector components :<br />

g v ], is the viscous flux tensor with the following<br />

⎡<br />

⎤ ⎡<br />

⎤<br />

0<br />

0<br />

f v =<br />

τ xx<br />

⎢<br />

⎣ τ<br />

⎥<br />

xy ⎦ , g v =<br />

τ yx<br />

⎢<br />

⎣ τ<br />

⎥<br />

yy ⎦ , (2.30)<br />

uτ xx + vτ xy + q x uτ yx + vτ yy + q y<br />

and τ is the two dimensional viscous stress tensor :<br />

τ =<br />

[<br />

τxx τ xy<br />

τ yx τ yy<br />

]<br />

. (2.31)


CHAPTER 2. GOVERNING EQUATIONS OF GAS DYNAMICS 45<br />

Assuming a Newtonian fluid, for which the Stokes hypothesis is valid, then τ<br />

is symmetric and a linear function <strong>of</strong> the velocity gradients:<br />

τ xx = − 2 µ∇ · u + 2µ∂u<br />

3 ∂x ,<br />

(2.32a)<br />

τ yy = − 2 µ∇ · u + 2µ∂v<br />

3 ∂y ,<br />

where µ is the dynamic viscosity <strong>of</strong> the fluid.<br />

(2.32b)<br />

( ∂u<br />

τ xy = τ yx = µ<br />

∂y + ∂v )<br />

. (2.32c)<br />

∂x<br />

The heat fluxes are modeled according to Fourier’s law:<br />

q x = −λ ∂T<br />

∂x , (2.33)<br />

q y = −λ ∂T<br />

∂y , (2.34)<br />

where the thermal conductivity coefficient λ is related to the dynamic viscosity and<br />

the specific heat at constant pressure c p by the non-dimensional Prandtl number:<br />

P r = µc p<br />

λ . (2.35)<br />

2.2.1 Mathematical character <strong>of</strong> the Navier-Stokes equations<br />

The presence <strong>of</strong> viscosity and heat conduction effects transforms the momentum<br />

and energy equations into second-order partial differential equations. This results<br />

in a hybrid system being parabolic-hyperbolic in time and space, but becoming<br />

elliptic-parabolic in space for steady state problems.


CHAPTER 2. GOVERNING EQUATIONS OF GAS DYNAMICS 46<br />

2.3 Non-dimensional form <strong>of</strong> the Navier-Stokes<br />

equations<br />

In order to study flows under different conditions around the same geometry and,<br />

moreover, in order to reduce the errors due to the finite precision <strong>of</strong> computers, a<br />

non-dimensionalization <strong>of</strong> the flow variables is performed. In Table 2.1 the reference<br />

quantities used are given, which were taken from [118].<br />

Variable Reference quantity<br />

ρ<br />

ρ ∞<br />

u, v C ∞<br />

E<br />

ρ ∞ C∞<br />

2<br />

p<br />

ρ ∞ C∞<br />

2<br />

x, y L<br />

T<br />

T ∞<br />

µ µ ∞<br />

λ<br />

λ ∞<br />

t<br />

L/C ∞<br />

Table 2.1: Reference quantities used for the non-dimensionalization <strong>of</strong> the Navier-<br />

Stokes equations.<br />

The dimensionless variables, denoted by an over-bar, are the following:<br />

ū, ¯v = u, v , ¯p = p , Ē = E ,<br />

C ∞ ρ ∞ C∞<br />

2 ρ ∞ C∞<br />

2<br />

¯x, ȳ = x, y<br />

L ,<br />

¯t = tC ∞<br />

L ,<br />

¯T =<br />

T<br />

T ∞<br />

, ¯µ = µ<br />

µ ∞<br />

.<br />

Through the non-dimensionalization <strong>of</strong> the equations the following reference numbers<br />

appear:<br />

ˆ Reynolds number Re = ρ∞C∞L<br />

µ ∞<br />

ˆ Prandtl number P r = µ∞cp<br />

λ ∞


CHAPTER 2. GOVERNING EQUATIONS OF GAS DYNAMICS 47<br />

ˆ Mach number Ma = U∞<br />

C ∞<br />

Note that the Reynolds number Re uses c ∞ and therefore Re based on the free<br />

stream velocity (which is the case for experimentally given Reynolds numbers) must<br />

be scaled by the Mach number Ma. The equation <strong>of</strong> state Eq. (2.6) becomes:<br />

¯p =<br />

¯ρ ¯T<br />

γ , (2.36)<br />

and the viscous stresses and heat fluxes are altered, given by:<br />

¯τ xx = − 2 2¯µ ∂ū<br />

¯µ∇ · ū +<br />

3Re Re ∂¯x ,<br />

(2.37a)<br />

¯τ yy = − 2 2¯µ ∂¯v<br />

¯µ∇ · ū +<br />

3Re Re ∂ȳ ,<br />

¯τ xy = ¯τ yx =<br />

¯µ Re<br />

(2.37b)<br />

( ∂ū<br />

∂ȳ + ∂¯v )<br />

, (2.37c)<br />

∂¯x<br />

¯µ ∂<br />

¯q x = −<br />

¯T<br />

ReP r(γ − 1) ∂¯x ,<br />

¯µ ∂<br />

¯q y = −<br />

¯T<br />

ReP r(γ − 1) ∂ȳ .<br />

(2.37d)<br />

(2.37e)<br />

The dynamic viscosity <strong>of</strong> the fluid is related to the temperature <strong>of</strong> the flow<br />

using Sutherland’s law:<br />

where S = 100.55K.<br />

¯µ = µ ( ) 3<br />

T<br />

2 T∞ + S<br />

=<br />

µ ∞ T ∞ T + S , (2.38)


CHAPTER 2. GOVERNING EQUATIONS OF GAS DYNAMICS 48<br />

2.4 Integral form <strong>of</strong> the compressible flow equations<br />

The integral form <strong>of</strong> the Navier-Stokes equations is the basis <strong>of</strong> the FV methods.<br />

Considering an arbitrary fixed control volume Ω as it is shown in Fig. 2.1 and<br />

integrating Eq. (2.29) over it, follows:<br />

n<br />

Ω<br />

∂Ω<br />

Figure 2.1: Control Volume (CV) Ω.<br />

∫ ∫<br />

∫<br />

∂<br />

UdΩ + ∇ · F i dΩ = ∇ · F v dΩ ⇒<br />

∂t Ω<br />

Ω<br />

Ω<br />

∫ ∮<br />

∮<br />

∂<br />

UdΩ + F i · ndS = F v · ndS, (2.39)<br />

∂t Ω<br />

∂Ω<br />

∂Ω<br />

where n is the outward normal vector at the boundary <strong>of</strong> the control volume. Neglecting<br />

the viscous flux tensor in Eq. (2.39), the integral form <strong>of</strong> the Euler equations<br />

is obtained:<br />

∫ ∮<br />

∂<br />

UdΩ + F i · ndS = 0. (2.40)<br />

∂t Ω<br />

∂Ω


Chapter 3<br />

Mesh quality measures for the FV<br />

method<br />

The discretization methods mentioned in the introduction <strong>of</strong> the thesis are all mesh<br />

based methods and their solution quality is affected by the underlying mesh over<br />

which they are applied, especially the FV and the FD method. Thus, field simulations<br />

that incorporate complex physics and geometries poses a formidable challenge<br />

to grid generation [77]. The distribution <strong>of</strong> points and elements can be quite non–<br />

uniform and coarse leading to inaccurate computations. The inaccuracy is <strong>of</strong>ten<br />

assessed after the simulation is performed causing repetitive grid generations and<br />

subsequent field simulations, so it would be desirable to have an assessment <strong>of</strong> the<br />

appropriateness (quality) <strong>of</strong> the mesh before performing the simulation. The presented<br />

study on the mesh quality focuses on the FV method as its use is currently<br />

widespread in CFD.<br />

Grid or mesh quality is affected by two primary factors; the local size <strong>of</strong><br />

the computational elements, and the uniformity <strong>of</strong> the spatial distribution <strong>of</strong> the<br />

points/elements. A primary method for solving the issue <strong>of</strong> coarse mesh resolution<br />

has been adaptive grid refinement [11, 70, 71, 76, 102]. While substantial work<br />

has been performed on this type <strong>of</strong> mesh adaptation, relatively less work has been<br />

devoted to study and improvement <strong>of</strong> the local distribution <strong>of</strong> the points, which<br />

49


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 50<br />

basically relates to the shape <strong>of</strong> the elements. Assessment <strong>of</strong> adequacy <strong>of</strong> local resolution<br />

and shape <strong>of</strong> the elements depends on the discretization error. It is imperative<br />

to derive measures <strong>of</strong> this error in the computation.<br />

Two broad categories <strong>of</strong> error indicators are: (i) a priori and (ii) a posteriori<br />

estimation. The two approaches are basically complimentary. A priori error<br />

evaluation can aid in mesh generation, while a posteriori can provide guidance to<br />

mesh adaptation techniques during the simulation. Up to now, works regarding<br />

the a priori grid quality assessment are based on geometric characteristics <strong>of</strong> the<br />

elements such as ratios <strong>of</strong> sizes <strong>of</strong> neighboring elements, as well as on element shape<br />

measures, such as angles and ratios <strong>of</strong> the radii <strong>of</strong> inscribed to prescribed circles<br />

[48, 52, 55, 80, 111]. In the FE method, the quality <strong>of</strong> a mesh is <strong>of</strong>ten given in terms<br />

<strong>of</strong> the element/mesh regularity. This type <strong>of</strong> approach has given measures that can<br />

be computed easily and are quite popular with practical applications.<br />

Truncation error (TE) analysis usually falls under the a posteriori estimation.<br />

The FD discretization method has <strong>of</strong>fered a vehicle for calculating the TE by<br />

performing a Taylor series expansions <strong>of</strong> the solution at the points forming the discretization<br />

stencil [132]. The complexity <strong>of</strong> the expressions has led research works to<br />

focus on simplified model field equations [8, 127, 154]. Nevertheless, those works expressed<br />

the strong dependence <strong>of</strong> the solution and the stability <strong>of</strong> the computations<br />

on the grid and its quality.<br />

Direct relations between TE and mesh distortion parameters, such as departure<br />

from orthogonality and uniformity have been reported in [66, 133]. Significantly less<br />

work exists for unstructured meshes [57] and for the FV method [72, 141]. However,<br />

despite the large number <strong>of</strong> studies on the relation on the TE to the numerical<br />

solution, there have been very few derivations <strong>of</strong> mesh quality measures based on it.<br />

A posteriori error estimation methods include the Richardson extrapolation<br />

[25, 26, 67, 134]. Use is made <strong>of</strong> two or more grids <strong>of</strong> the same domain with the<br />

difference in the yielded solution <strong>of</strong>fering a measure <strong>of</strong> the local error distribution.<br />

Generating and solving on multiple meshes can be quite difficult for practical applications.


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 51<br />

Another a posteriori method evaluates the error indirectly via tracking <strong>of</strong> the<br />

numerical solution variation [78]. Local field features, such as boundary layers, shock<br />

waves and vortices are detected using sensors that are based on variations <strong>of</strong> the<br />

computed field parameters, such as the pressure and velocity. The assumption here<br />

is that the discretization error is large where the solution variations are large. The<br />

method has been primarily used for guiding grid adaptation. It is not a method<br />

that can directly yield practical assessment <strong>of</strong> grid quality, taking also into account<br />

that it necessitates expensive computations for large meshes and/or complex fields.<br />

Another school <strong>of</strong> a posteriori error estimation work employs FE discretization<br />

and derives analytic expressions <strong>of</strong> error bounds [3, 106, 116, 117]. Various model<br />

equations have been utilized in order to provide the estimates. Although, the error<br />

bounds do not yield grid quality measures directly, there is potential for such use.<br />

Again, knowledge <strong>of</strong> the solution field is needed.<br />

The goal <strong>of</strong> the mesh induced error study is the derivation <strong>of</strong> a priori quality<br />

measures for mixed-type element meshes for the FV method. The most fundamental<br />

computation is that <strong>of</strong> first-order derivatives. The first derivative is common not<br />

only to fluid flow governing equations, but also to other field equations. The relative<br />

reduced complexity <strong>of</strong> the related mathematics and its commonality led to its choice<br />

in the present work. The FV method will be the method for the mesh induced error<br />

study, as the FV method is currently the most wide spread method in computational<br />

fluid dynamics and moreover, it is the first-order accurate version <strong>of</strong> the DG<br />

discretization. A node-based FV method is chosen as the vehicle for derivation <strong>of</strong><br />

quality measures, since it is known to be sensitive to mesh non-uniformity in terms<br />

<strong>of</strong> its accuracy. The complete expressions for the TE will be presented for meshes<br />

that consist <strong>of</strong> structured, unstructured, or both types <strong>of</strong> elements. The complexity<br />

<strong>of</strong> the analytic expressions lead to study <strong>of</strong> the two–dimensional case.


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 52<br />

3.1 Gradient computation in the FV method<br />

A common node based FV evaluation <strong>of</strong> the first-order spatial derivatives employs<br />

the median dual area around a grid point as Fig. 3.1 depicts [15, 63]. This quite<br />

common FV evaluation is chosen to study quality measures. It is expected to reveal<br />

distortions <strong>of</strong> the mesh as it corresponds to a central differencing, and thus, deemed<br />

sufficient for this study.<br />

0 u e<br />

∆x e<br />

∆y e<br />

i-1<br />

e, 1<br />

i<br />

e, 2<br />

i+1<br />

Figure 3.1: Median dual surface for evaluating first-order derivatives at grid point<br />

0 via the FV method.<br />

The evaluation <strong>of</strong> the gradient ∇u is accomplished via the following contour<br />

integration:<br />

∇u ≈ 1 S<br />

∮<br />

∂l<br />

udl ≈ 1 ∑<br />

u e (∆y e î − ∆x e ĵ) ≡ ∇ h u, (3.1)<br />

S<br />

e<br />

where S is the dual area, u e is the field value at the middle <strong>of</strong> each edge <strong>of</strong> the dual<br />

contour , while ∆x e and ∆y e are the projections in the x and y direction <strong>of</strong> the<br />

edges <strong>of</strong> the dual contour joining the centroids <strong>of</strong> each element sharing point 0 with


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 53<br />

the middle <strong>of</strong> the grid edges sharing this point . This contour is termed median<br />

dual [15]. Other choices <strong>of</strong> dual surfaces include the surface formed by joining the<br />

centroids <strong>of</strong> the elements sharing point 0 (centroid dual), as well as the surface <strong>of</strong><br />

the union <strong>of</strong> those elements.<br />

Since the goal is deriving a quality measure the specific choice is not important.<br />

A node-based evaluation <strong>of</strong> the derivatives is considered, which will reveal the<br />

inaccuracy due to the distorted elements in a relatively straightforward manner.<br />

3.2 The general form <strong>of</strong> the truncation error for<br />

the node based FV method<br />

The TE in the computation <strong>of</strong> the gradient is defined as:<br />

E(x, y) = ∇ h u(x, y) − ∇u(x, y), (3.2)<br />

with ∇ h u and ∇u being the numerical and the analytical values <strong>of</strong> the gradient,<br />

respectively. The analysis is similar in the x and y directions and employs Taylor<br />

series expansions for u h at the required locations around point 0. The expansions are<br />

substituted into the TE definition <strong>of</strong> Eq. (3.2). The amount <strong>of</strong> operations involved<br />

is very large. This must have been one <strong>of</strong> the main reasons for not seeing complete<br />

TE analysis in previous works. In the present work the hurdle is overcome via use<br />

<strong>of</strong> the symbolic mathematics capability <strong>of</strong> Matlab [101].<br />

The TE terms are grouped according to the following general form given for<br />

the case <strong>of</strong> evaluating the derivative in the x direction (u h x):<br />

E x =e x xu x + e x yu y + e x xyu xy + e x xxu xx + e x yyu yy +<br />

e x xxxu xxx + e x yyyu yyy + e x xyyu xyy + e x xxyu xxy + ...<br />

(3.3)<br />

The terms with the symbol e are functions <strong>of</strong> the metrics <strong>of</strong> the mesh and will be


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 54<br />

the link for making the connection <strong>of</strong> the TE with mesh distortion. They will be<br />

termed error coefficients (EC). Their analytic expressions are grouped according to<br />

the order in terms <strong>of</strong> the local mesh size h.<br />

The first two terms are involved in the consistency checks <strong>of</strong> the numerical<br />

approximation:<br />

e x x = 1<br />

2S<br />

e x y = 1<br />

2S<br />

n∑<br />

(∆x e,1 + ∆x e,2 )∆y e − 1,<br />

e=1<br />

n∑<br />

(∆y e,1 + ∆y e,2 )∆y e .<br />

e=1<br />

(3.4a)<br />

(3.4b)<br />

The grid metrics involved in the above expressions are defined as follows:<br />

(∆x e,k ) j (∆y e,k ) l = 1<br />

N e,k<br />

N e,k<br />

∑<br />

(x m | e,k − x 0 ) j (y m | e,k − y 0 ) l ,<br />

m=1<br />

(3.5)<br />

∆y e = y e,2 − y e,1 ,<br />

where x m | e,k , y m | e,k and N e,k are the coordinates and the number <strong>of</strong> nodes participating<br />

in the averaging <strong>of</strong> u at the dual vertex e, k (k = 1, 2). Generally, if a<br />

median dual vertex coincides with the middle point <strong>of</strong> an edge sharing point 0, then<br />

N e,k = 2, and if it coincides with the center <strong>of</strong> a quadrilateral or a triangle then<br />

N e,k = 4, or N e,k = 3, respectively. In Eq. (3.5) it is observed that the metric terms<br />

∆x e,1 , ∆x e,2 , ∆y e,1 and ∆y e,2 are O(h) each.<br />

In the above expressions, n is the number <strong>of</strong> the dual edges, S is the dual area<br />

and it is O(h 2 ), while e, 1 and e, 2 denote the end points <strong>of</strong> each edge <strong>of</strong> the dual<br />

contour, as shown in Fig. 3.1.<br />

The next three terms are first-order:


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 55<br />

e x xy = 1<br />

2S<br />

e x xx = 1<br />

2S2!<br />

e x yy = 1<br />

2S2!<br />

n∑<br />

(∆x e,1 ∆y e,1 + ∆x e,2 ∆y e,2 )∆y e ,<br />

e=1<br />

n∑<br />

[(∆x e,1 ) 2 + (∆x e,2 ) 2 ]∆y e ,<br />

e=1<br />

n∑<br />

[(∆y e,1 ) 2 + (∆y e,2 ) 2 ]∆y e .<br />

e=1<br />

(3.6a)<br />

(3.6b)<br />

(3.6c)<br />

For the purpose <strong>of</strong> deriving quality measures, the second-order derivative terms are<br />

retained as well:<br />

e x xxx = 1<br />

2S3!<br />

e x yyy = 1<br />

2S3!<br />

e x xyy = 1<br />

2S2!<br />

e x xxy = 1<br />

2S2!<br />

n∑<br />

[(∆x e,1 ) 3 + (∆x e,2 ) 3 ]∆y e ,<br />

e=1<br />

n∑<br />

[(∆y e,1 ) 3 + (∆y e,2 ) 3 ]∆y e ,<br />

e=1<br />

n∑<br />

[∆x e,1 (∆y e,1 ) 2 + ∆x e,2 (∆y e,2 ) 2 ]∆y e ,<br />

e=1<br />

n∑<br />

[(∆x e,1 ) 2 ∆y e,1 + (∆x e,2 ) 2 ∆y e,2 ]∆y e .<br />

e=1<br />

(3.7a)<br />

(3.7b)<br />

(3.7c)<br />

(3.7d)<br />

When the projections ∆x e and ∆y e are symmetrical relative to point 0, the<br />

computation becomes <strong>of</strong> second-order and the TE expression reduces to the following:


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 56<br />

E x = e x xxxu xxx + e x xyyu xyy ,<br />

E y = e y yyyu yyy + e y xxyu xxy .<br />

(3.8)<br />

Meshes that give a TE form given in Eq. (3.8) will be termed ideal meshes. Orthogonal<br />

and equally spaced in each direction quadrilaterals form an ideal mesh. The<br />

equivalent with triangles is an unstructured mesh with equal-length projections <strong>of</strong><br />

their dual edges in the x and y directions. This definition will be used subsequently<br />

to define a quality measure.<br />

Similar expressions for the EC regarding evaluation <strong>of</strong> the derivative in the y<br />

direction hold. The projection ∆y e is replaced with ∆x e and a minus sign is placed<br />

in front <strong>of</strong> the sums. The EC e y x and e y y are given below:<br />

e y x = − 1<br />

2S<br />

e y y = − 1<br />

2S<br />

n∑<br />

(∆x e,1 + ∆x e,2 )∆x e ,<br />

e=1<br />

n∑<br />

(∆y e,1 + ∆y e,2 )∆x e − 1.<br />

e=1<br />

(3.9a)<br />

(3.9b)<br />

3.3 Consistency condition for the node based FV<br />

method<br />

A consistent discretization <strong>of</strong> the gradient requires the TE to vanish as the mesh<br />

size approaches zero. For zero mesh size all the EC become equal to zero, except<br />

for e x x, e x y in Eq. (3.4) and e y x, e y y in Eq. (3.9), and the TE in Eq. (3.2) becomes:


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 57<br />

E x = e x xu x + e x yu y ,<br />

E y = e y yu y + e y xu x .<br />

(3.10)<br />

It is noted that the EC e x x, e y x, e x y, e y y remain unchanged for fixed–shape mesh elements<br />

even if the local mesh size is changing. For a consistent discretization, it is required<br />

that:<br />

e x x = e x y = e y x = e y y = 0. (3.11)<br />

Following Eq. (3.11) the expressions <strong>of</strong> the EC e x x in Eq. (3.4) and e y y in Eq.<br />

(3.9) show that in order for them to vanish on a general mesh, the area evaluation<br />

<strong>of</strong> the following two expressions must yield the same result.<br />

S x = 1 n∑<br />

(∆x e,1 + ∆x e,2 )∆y e ,<br />

2<br />

S y = − 1 2<br />

e=1<br />

n∑<br />

(∆y e,1 + ∆y e,2 )∆x e .<br />

e=1<br />

(3.12)<br />

Using the median dual surface, from Eq. (3.12) it results that the two forms<br />

for evaluating the dual area produce the same result. Specifically, their difference is<br />

the following:<br />

∑<br />

[x e,2 y e,2 − x e,1 y e,1 − 2x 0 (y e,2 − y e,1 )] = 0, (3.13)<br />

e<br />

and so e x x = e y y = 0 on a mixed-type element mesh. Carrying out the algebra in the<br />

expression <strong>of</strong> the EC e x y in Eq. (3.4) using the median dual results:


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 58<br />

e x y = 1<br />

2S<br />

Analogously, the EC e y x in Eq.<br />

∑<br />

[ye,2 2 − ye,1 2 − 2y 0 (y e,2 − y e,1 )] = 0. (3.14)<br />

e<br />

(3.9) results to be always zero on a mixed-type<br />

element mesh. So, the median dual surface leads to a consistent approximation <strong>of</strong><br />

∇u.<br />

Using the centroid dual, from Eq.<br />

(3.12) it results that the two forms for<br />

evaluating the dual area do not produce the same result on structured meshes.<br />

Their difference considering a structured mesh is:<br />

1<br />

8<br />

∑<br />

[x i (y i+1 − y i−1 ) + y i (x i+1 − x i−1 )] ≠ 0, (3.15)<br />

i<br />

and so e x x ≠ 0, e y y ≠ 0 on a structured mesh. Carrying out the algebra in the<br />

expression <strong>of</strong> the EC e x y in Eq. (3.4) using the centroid dual it gives:<br />

e x y = 1<br />

8S<br />

∑<br />

[y i (y i+1 − y i−1 )] ≠ 0. (3.16)<br />

i<br />

Analogously, the EC e y x in Eq. (3.9) is not always zero on structured meshes. So,<br />

the consistency <strong>of</strong> the centroid dual depends on the geometry <strong>of</strong> a structured mesh<br />

in the approximation <strong>of</strong> ∇u.<br />

Using the centroid dual, from Eq.<br />

(3.12) it results that the two forms for<br />

evaluating the dual area produce the same result on unstructured meshes.<br />

difference <strong>of</strong> the two forms in Eq. (3.12) considering an unstructured mesh is:<br />

1<br />

6<br />

The<br />

∑<br />

[x i (y i+1 − y i−1 ) + x i (y i+1 − y i−1 )] = 0, (3.17)<br />

i<br />

and so e x x = e y y = 0 on unstructured meshes. Carrying out the algebra in the<br />

expression <strong>of</strong> the EC e x y in Eq. (3.4) using the centroid dual it gives:<br />

e x y = 1<br />

6S<br />

∑<br />

[y i (y i+1 − y i−1 )] = 0. (3.18)<br />

i


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 59<br />

Analogously, the EC e y x in Eq. (3.9) is always zero on unstructured meshes. So, the<br />

centroid dual gives a consistent approximation <strong>of</strong> ∇u on unstructured meshes.<br />

At mixed-type element mesh interfaces it is difficult to check analytically the<br />

consistency <strong>of</strong> the approximation <strong>of</strong> ∇u using the centroid dual. Despite this fact,<br />

it is anticipated that the computation will be inconsistent due to the inconsistency<br />

<strong>of</strong> the centroid dual expressions on structured meshes.<br />

3.4 Verification <strong>of</strong> the error coefficients analytic<br />

expressions<br />

It is <strong>of</strong> importance to check the complex expression <strong>of</strong> the TE using analytic field<br />

functions u(x, y). The TE is computed by substituting the analytic values <strong>of</strong> the<br />

spatial derivatives and the values <strong>of</strong> the EC in Eq. (3.3). This will be termed<br />

the analytic TE. The TE is also computed directly by subtracting the analytic<br />

expression for ∇u from the FV expression ∇ h u (Eq. (3.2)). This second approach<br />

will be termed numerical TE.<br />

Let us consider a field function u(x, y), which resembles a boundary layer type<br />

<strong>of</strong> flow field with a linear variation in the streamwise direction:<br />

u(x, y) = xy 2 . (3.19)<br />

The differences between the analytic and the numerical TE values are calculated for<br />

the mixed-type element mesh <strong>of</strong> Fig. 3.2 using both the median and the centroid<br />

dual. A feature <strong>of</strong> this mesh is a local lateral displacement in the vicinity <strong>of</strong> an<br />

interface point.<br />

For the median dual the distribution <strong>of</strong> the analytic and the numerical TE is shown<br />

in Fig. 3.3. It is observed that the derived expressions are coincident within plotting<br />

accuracy.<br />

Check <strong>of</strong> consistency “experimentally” via successive reductions <strong>of</strong> the size <strong>of</strong>


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 60<br />

Figure 3.2: Mixed-type element channel mesh.<br />

the mesh elements may be impractical for large meshes. An indirect check <strong>of</strong> the<br />

consistency is proposed here by computing the difference between the analytic and<br />

numerical TE, without including in the analytic expression the error coefficients<br />

“responsible” for consistency (e x x and e x y). When including these terms, which are<br />

independent <strong>of</strong> the local mesh size, the difference will be zero within machine accuracy.<br />

If it is not, as the case is for the centroid dual discretization (see Fig. 3.4), a<br />

local inconsistency is revealed. The “spikes” in Fig. 3.4 correspond to points in the<br />

vicinity <strong>of</strong> the locally distorted mesh interface. The same inconsistency was revealed<br />

for the case <strong>of</strong> a distorted structured mesh.<br />

Similar results were observed using other field functions, as well. This provides<br />

assurance that the derived complex expressions are correct. A further check <strong>of</strong> their<br />

validity, even though indirect, will be provided when checking the subsequently<br />

derived quality measures.<br />

3.5 Elementary Types <strong>of</strong> Mesh Distortion<br />

The general mesh depicted in Fig. 3.1 exhibits skewness, stretching, <strong>of</strong>fset <strong>of</strong> point<br />

0 from the center <strong>of</strong> the control surface, as well as interfaces between structured


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 61<br />

0.01<br />

0.005<br />

Error<br />

0<br />

-0.005<br />

-0.01<br />

0 20 40 60 80 100 120 140 160 180 200 220<br />

Node<br />

Figure 3.3: Distribution <strong>of</strong> the analytic (✷) and the numerical (✸) TE for the field<br />

u(x, y) = xy 2 on the mesh <strong>of</strong> Fig. 3.2 using the median dual.<br />

and unstructured elements. It is important in the context <strong>of</strong> the present work to<br />

“isolate” each type <strong>of</strong> mesh “defect” and relate it directly to the TE. This is very<br />

important for future work on improving the mesh during or immediately after its<br />

generation. Distortion types are easier to define on a quadrilateral mesh, compared<br />

to a triangular for which the <strong>of</strong>fset <strong>of</strong> point 0 from the dual centroid is a reasonable<br />

measure. Regarding the interfaces between quadrilateral and triangular elements,<br />

three distinct types were identified and studied. It should be noted that skewness,<br />

stretching and center point <strong>of</strong>fset are distortions that are not independent from each<br />

other. Application <strong>of</strong> one <strong>of</strong> them to an ideal mesh causes appearance <strong>of</strong> the rest.<br />

Nevertheless, this categorization <strong>of</strong>fers a clear way to express the TE as a direct<br />

function <strong>of</strong> them and to come up with ways to reduce those distortions. It should<br />

be noted that the present work considers interior point configurations. Boundary<br />

“central” points (node 0 in Fig. 3.1) are not examined as the application <strong>of</strong> boundary<br />

conditions <strong>of</strong>ten “discards” the local discretization.


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 62<br />

0<br />

Difference in Error<br />

-0.001<br />

-0.002<br />

0 20 40 60 80 100 120 140 160 180 200 220<br />

Node<br />

Figure 3.4: Difference between the analytic and the numerical TE for the field<br />

u(x, y) = xy 2 on the mesh <strong>of</strong> Fig. 3.2 using the centroid dual.<br />

3.6 Stretching<br />

Stretching is a common type <strong>of</strong> deviation from a uniformly–spaced mesh and it is<br />

readily defined for structured grids. In order to compare quadrilaterals and triangles<br />

in terms <strong>of</strong> sensitivity to stretching, the stretching type <strong>of</strong> mesh distortion is defined<br />

in a similar way as Fig. 3.5 depicts. In the quadrilateral case, the stretching is<br />

considered as the displacement <strong>of</strong> the edges sharing the center point 0 quantified<br />

by the stretching factors d x and d y . These edges are displaced in a structured<br />

way retaining their orientation, while stretching in the triangle case is defined via<br />

displacing the center node according to Fig. 3.5(b).


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 63<br />

dx∆x<br />

•<br />

∆y<br />

0<br />

dy∆y<br />

∆y<br />

∆x<br />

(a)<br />

∆x<br />

d x h<br />

•<br />

d y h<br />

h<br />

0<br />

(b)<br />

Figure 3.5: Stretching for (a) structured and (b) unstructured meshes defined to<br />

facilitate comparison <strong>of</strong> accuracy degradation on them.<br />

3.6.1 Skewness and Shearing<br />

The four edges sharing the center point 0 in a quadrilateral grid usually meet forming<br />

angles that deviate from 180 ◦ as Fig. 3.6 illustrates. This type <strong>of</strong> distortion is<br />

typically called skewness. The angle ω quantifies the degree <strong>of</strong> skewness and ranges<br />

from 0 ◦ to 90 ◦ . Shearing is quite different from skewness. It preserves the 180 ◦<br />

angle between the edges and deviates the mesh from being orthogonal (Fig. 3.7).<br />

The angle φ quantifies this deviation and ranges from 0 ◦ to 90 ◦ . It will be seen that<br />

shearing and skewness affect the accuracy very differently.


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 64<br />

0 ω<br />

∆x<br />

Figure 3.6: Skewed mesh, depicting the deviation angle ω from the 180 ◦ angle<br />

between one <strong>of</strong> the two pairs <strong>of</strong> edges sharing point 0 (ω ∈ [0 ◦ , 90 ◦ ]). The lengths <strong>of</strong><br />

the edges are equal.<br />

3.6.2 Mesh rotation<br />

Non–alignment <strong>of</strong> a structured mesh to a specific direction dictated by the form<br />

<strong>of</strong> the field is another type <strong>of</strong> distortion examined. The quadrilateral elements<br />

are rotated with respect to the x–axis <strong>of</strong> the coordinate system by an angle θ as<br />

illustrated in Fig. 3.8. This angle varies from 0 ◦ to 45 ◦ .<br />

0<br />

φ<br />

∆x<br />

Figure 3.7: Sheared mesh, depicting displacement <strong>of</strong> the edges causing deviation<br />

from orthogonality expressed by the angle φ (φ ∈ [0 ◦ , 90 ◦ ]). The lengths <strong>of</strong> the<br />

edges are equal.


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 65<br />

∆y<br />

y<br />

∆y<br />

0<br />

θ<br />

x<br />

∆x<br />

∆x<br />

Figure 3.8: Rotated structured mesh depicting non alignment <strong>of</strong> the edges with the<br />

axes <strong>of</strong> the global system (∆x ≠ ∆y and θ ∈ [0 ◦ , 45 ◦ ]).<br />

3.6.3 Mixed-type element mesh interfaces<br />

Change in the topology <strong>of</strong> the elements creates special interfaces which require<br />

examination in terms <strong>of</strong> the local accuracy <strong>of</strong> the discretization [75]. The elements<br />

forming the interface (quadrilaterals and triangles) are considered to have the same<br />

size so that the effect <strong>of</strong> the changing topology is isolated and studied. Three cases<br />

<strong>of</strong> such interfaces are identified in two dimensions and are illustrated in Fig. 3.9.<br />

ψ 2 ψ 1<br />

h<br />

h<br />

h<br />

0<br />

0<br />

0<br />

h<br />

h<br />

h<br />

h<br />

(a)<br />

h<br />

(b)<br />

h<br />

(c)<br />

Figure 3.9: Common mixed-type element mesh interfaces.


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 66<br />

3.7 Direct relation between truncation error and<br />

mesh distortion<br />

The general expressions <strong>of</strong> the error coefficients (EC) given in Eqs. (3.6) and (3.7)<br />

can yield simpler expressions directly related to each type <strong>of</strong> mesh distortion presented<br />

here. This is very important to subsequent work on improving the grid.<br />

Table 3.1 gives the expressions <strong>of</strong> the EC for a mesh exhibiting only one type<br />

<strong>of</strong> distortion for each case (a row <strong>of</strong> the table). The related parameters d x , d y , ω<br />

are present in the expressions. For clarity <strong>of</strong> the table, the terms related to the<br />

higher order EC are not given, and it is just noted if they are zero or not. It is<br />

observed that stretching and skewness reduce the formal second order <strong>of</strong> the FV<br />

discretization to first. The error is proportional to the stretching parameters d x and<br />

d y . In the case <strong>of</strong> skewness, the relationship with ω is not linear. Finally, for the<br />

case <strong>of</strong> shearing and non–alignment <strong>of</strong> a structured mesh, no reduction <strong>of</strong> the order<br />

<strong>of</strong> accuracy occurs. However, the structure <strong>of</strong> the TE changes (terms e x xxx, e x yyy, e x xyy<br />

and e x xxy) but the discretization remains second order.<br />

Mesh e x xy e x xx e x yy e x xxx e x yyy e x xyy e x xxy<br />

Fig. 3.5(a)<br />

1<br />

2 d y∆y d x ∆x 0 ≠ 0 0 ≠ 0 ≠ 0<br />

Fig. 3.5(b) d y h d x h 0 ≠ 0 0 ≠ 0 ≠ 0<br />

Fig. 3.6<br />

3 sin (2ω)∆x<br />

8 cos (ω)+8<br />

[cos (ω)−1]∆x<br />

2<br />

[1−cos (ω)]∆x<br />

4<br />

≠ 0 ≠ 0 ≠ 0 ≠ 0<br />

Fig. 3.7 0 0 0 ≠ 0 0 ≠ 0 ≠ 0<br />

Fig. 3.8 0 0 0 ≠ 0 ≠ 0 ≠ 0 ≠ 0<br />

Table 3.1: Expressions for the error coefficients for each type <strong>of</strong> mesh distortion.<br />

The EC corresponding to evaluation <strong>of</strong> the derivative u y <strong>of</strong> the three types <strong>of</strong>


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 67<br />

mixed-type element mesh interfaces are presented in Table 3.2. For the first type<br />

<strong>of</strong> interface, it is considered that ψ 1 = ψ 2 . It is observed that mesh interfaces yield<br />

first order accuracy for the evaluation <strong>of</strong> the derivative u y .<br />

Mesh e y xy e y xx e y yy e y xxx e y yyy e y xyy e y xxy<br />

Fig. 3.9(a) 0 − h 48<br />

0 ≠ 0 ≠ 0 ≠ 0 ≠ 0<br />

Fig. 3.9(b) 0 0 − h 10<br />

0 ≠ 0 ≠ 0 ≠ 0<br />

Fig. 3.9(c) − h 6<br />

0 0 0 ≠ 0 ≠ 0 ≠ 0<br />

Table 3.2: Expressions for the error coefficients for the mixed-type element mesh<br />

interfaces.<br />

However, it is possible to relocate node 0, so that the order <strong>of</strong> accuracy <strong>of</strong> the u y<br />

evaluation is improved, but only for the mesh interfaces <strong>of</strong> Figs. 3.9(b),(c). In Table<br />

3.3 the displacements <strong>of</strong> point 0 for improving the accuracy <strong>of</strong> the u y evaluation are<br />

given. It should be noted that the results <strong>of</strong> Table 3.3 concern interfaces with equal<br />

“heights” (h) <strong>of</strong> the elements.<br />

Mesh ∆x ∆y<br />

Fig. 3.9(a) non existent non existent<br />

Fig. 3.9(b) 0 2(−10 + 3 √ 11)h<br />

Fig. 3.9(c) − h 6<br />

0<br />

Table 3.3: Central interface point displacements for improving the accuracy <strong>of</strong> the<br />

u y derivative evaluation for the mixed-type element mesh interfaces.


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 68<br />

3.8 Definition <strong>of</strong> an appropriate index <strong>of</strong> mesh<br />

distortion<br />

The focus <strong>of</strong> the present work is on a priori evaluation <strong>of</strong> a grid with regards to<br />

the shape and topology <strong>of</strong> its elements. The other aspect <strong>of</strong> grid quality, namely<br />

the local resolution is not addressed here. Therefore, the defined measure (index) <strong>of</strong><br />

mesh distortion should be independent <strong>of</strong> the local size <strong>of</strong> the mesh. Other properties<br />

that this index should have include:<br />

1. simple mathematical form,<br />

2. direct relation with the TE, as well as the mesh distortion parameters,<br />

3. ability to capture distortions in any direction, and<br />

4. ability to detect relatively small distortions in the presence <strong>of</strong> larger ones.<br />

3.8.1 Normalized error coefficients<br />

The EC <strong>of</strong> Eqs. (3.6) and (3.7) are divided by the appropriate power <strong>of</strong> a characteristic<br />

local length scale L. Care must be taken in order to account for directionally-sized<br />

(high aspect ratio) local mesh elements. This implies appropriate use <strong>of</strong> two length<br />

scales (L x ,L y ) expressing the local size in the x and y direction, respectively.<br />

The “normalized” EC <strong>of</strong> Eqs. (3.6) and (3.7) are defined as :<br />

e x xx = ex xx<br />

L x<br />

,<br />

e x yy = ex yy<br />

L y<br />

,<br />

e x xy = ex xy<br />

L y<br />

, (3.20)<br />

e x xxx = ex xxx<br />

L 2 x<br />

, e x yyy = ex yyy<br />

L 2 y<br />

, e x xxy = ex xxy<br />

L 2 y<br />

, e x xyy = ex xyy<br />

. (3.21)<br />

L 2 y<br />

The denumerators in the above definitions have been chosen based on the mesh<br />

metrics appearing in the EC expressions and the order <strong>of</strong> each EC.


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 69<br />

Careful consideration is needed to define the characteristic local length. It is<br />

incorrect to define a single length for every direction, as then, on meshes with cells<br />

<strong>of</strong> high aspect ratio, the normalized EC would either be overestimated or underestimated.<br />

The present work uses an approach that resembles the FV evaluation <strong>of</strong><br />

the derivative. It defines the following grid functions :<br />

f x = ∆x e,k |∆x e,k |, f y = ∆y e,k |∆y e,k |, k = 1, 2. (3.22)<br />

Then, the lengths L x ,L y are computed in a similar manner to Eq. (3.1) with u being<br />

replaced by f x for the x–component and by f y for the y–component:<br />

L x = 1 ∑<br />

f x,e ∆y e ,<br />

S<br />

e<br />

L y = − 1 ∑<br />

f y,e ∆x e (3.23)<br />

S<br />

e<br />

It should be noted that for a uniform structured mesh with element sizes ∆x and<br />

∆y, the characteristic lengths are:<br />

L x = ∆x and L y = ∆y.<br />

For the case <strong>of</strong> a uniform (honeycomb shape) triangular mesh <strong>of</strong> edge size h, the<br />

local lengths are:<br />

L x = 35h<br />

48<br />

and<br />

L y = 7√ 3h<br />

16 .<br />

Also, for the examined mixed-type element mesh interfaces the characteristic local<br />

lengths are given in Table 3.4.<br />

3.8.2 Mesh quality index<br />

The normalized EC are grouped to yield a single number that characterizes the<br />

quality <strong>of</strong> the grid. Two such groupings are being presented and evaluated. The<br />

first is expressed via use <strong>of</strong> the first-order error coefficients:


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 70<br />

Mesh L x L y<br />

Mesh interface <strong>of</strong> Fig. 3.9(a)<br />

11<br />

12 h h<br />

Mesh interface <strong>of</strong> Fig. 3.9(b) h h<br />

Mesh interface <strong>of</strong> Fig. 3.9(c) h h<br />

Table 3.4: Characteristic local lengths for the mixed-type element mesh interfaces.<br />

Q x = |e x xx| + |e x yy| + |e x xy|,<br />

(3.24a)<br />

Q y = |e y xx| + |e y yy| + |e y xy|,<br />

(3.24b)<br />

where the first expression regards the u x derivative evaluation, and the second concerns<br />

the u y computation. On meshes which yield second-order accuracy in the<br />

evaluation <strong>of</strong> the derivatives u x and u y , both Q x and Q y are equal to zero.<br />

The second index that is presented expresses the deviation <strong>of</strong> a mesh from<br />

being ideal via the ratio <strong>of</strong> a form <strong>of</strong> the EC appearing on a general distorted mesh<br />

to a form <strong>of</strong> those coefficients for an ideal mesh. Specifically:<br />

q x = |ex xx| + |e x yy| + |e x xy| + |e x xxx| + |e x yyy| + |e x xyy| + |e x xxy|<br />

, (3.25a)<br />

|e x xxx| + |e x xyy|<br />

q y = |ey xx| + |e y yy| + |e y xy| + |e y xxx| + |e y yyy| + |e y xyy| + |e y xxy|<br />

. (3.25b)<br />

|e y xxx| + |e y xyy|<br />

Index q expresses the deviation <strong>of</strong> the mesh locally from being ideal. It should be


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 71<br />

noted that q x and q y are equal to one for an ideal mesh, and have values greater<br />

than one for a general distorted grid.<br />

It is interesting to examine the values <strong>of</strong> the indices for a structured mesh with<br />

high aspect ratio elements that is typical <strong>of</strong> boundary layer regions and illustrated<br />

in Fig. 3.10.<br />

0<br />

2∆y<br />

2∆x<br />

Figure 3.10: High aspect ratio quadrilaterals typical <strong>of</strong> a structured mesh in a<br />

boundary layer region.<br />

The relevant computation here is that <strong>of</strong> the derivative u y . The relevant EC<br />

in Eq. (3.8) for this case are:<br />

e y xxy = (∆x)2<br />

8<br />

, e y yyy = (∆y)2 ,<br />

6<br />

which yields the mesh quality index Q y to be zero and the q y to be one, which are<br />

the best possible quality index values.<br />

The experiment is repeated employing the triangular mesh created by subdividing<br />

the quadrilaterals <strong>of</strong> Fig. 3.10 along their diagonals. The corresponding error<br />

coefficients are:<br />

e y yyy = (∆y)2<br />

6<br />

, e y xyy = − ∆x∆y<br />

6<br />

, e y xxy = (∆x)2 .<br />

6<br />

It is observed that the order <strong>of</strong> accuracy remains second (Q y = 0). However, more<br />

non–zero EC appear and the value <strong>of</strong> q y is not that <strong>of</strong> the ideal mesh, but is greater<br />

than one.


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 72<br />

3.8.3 Analytic expressions <strong>of</strong> the quality index Q for each<br />

type <strong>of</strong> elementary mesh distortion<br />

The index Q has a simpler mathematical form and thus amenable to a direct relation<br />

to each type <strong>of</strong> mesh distortion. These expressions are given in Table 3.5. The<br />

equations for the stretched unstructured and the skewed structured mesh cases are<br />

given in Appendix A.<br />

Mesh<br />

Q<br />

Stretched structured [Fig. 3.5(a)]<br />

Stretched unstructured [Fig. 3.5(b)]<br />

Skewed [Fig. 3.6]<br />

Q x = | dx | + |<br />

1+d 2 x<br />

Q x (Eq. A-1)<br />

Q x (Eq. A-2)<br />

d y<br />

2+2d 2 y<br />

|<br />

Sheared [Fig. 3.7] Q x , Q y = 0<br />

Non–aligned structured [Fig. 3.8] Q x , Q y = 0<br />

Mesh interface [Fig. 3.9(a)] Q y = 1<br />

44<br />

Mesh interface [Fig. 3.9(b)] Q y = 144<br />

533<br />

Mesh interface [Fig. 3.9(c)] Q y = 1 6<br />

Table 3.5:<br />

distortion.<br />

Mesh quality index Q expressions for the elementary types <strong>of</strong> mesh<br />

It is observed that, the sheared and rotated (non–aligned) structured meshes do not<br />

exhibit reduction <strong>of</strong> accuracy and thus the indices are equal to zero. The unstructured<br />

mesh exhibits higher sensitivity in mesh distortion as it is depicted graphically<br />

in Fig. 3.11.<br />

The variation <strong>of</strong> the index with respect to the skewness angle is given in the Appendix<br />

A, and it is shown graphically in Fig. 3.12. A peak in the degradation <strong>of</strong>


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 73<br />

0.5<br />

0.4<br />

0.3<br />

Q x<br />

0.2<br />

0.1<br />

Unstructured<br />

Structured<br />

0<br />

0 0.1 0.2 0.3 0.4 0.5<br />

d x<br />

Figure 3.11: Mesh quality index Q x for the stretched unstructured mesh (◦) and the<br />

stretched structured mesh (□) vs. displacement factor d x (d y = 0).<br />

accuracy is observed at about 80 ◦ . It is observed that the normalized error coefficient<br />

e x xy reduces as the skewness angle approaches 90 ◦ , which leads to the reduction <strong>of</strong><br />

Q x in the same region in the graph.<br />

Finally, the influence <strong>of</strong> the mesh interface angle ψ 2 on degradation <strong>of</strong> accuracy<br />

is shown in the graph <strong>of</strong> Fig. 3.13. This concerns the mixed-type element<br />

mesh interface shown in Fig. 3.9(a). The curve is symmetric around the value <strong>of</strong><br />

tan −1 (2) ≈ 63 ◦ for which Q y is minimum.<br />

3.8.4 Calibration <strong>of</strong> the mesh quality index Q<br />

Any practical use <strong>of</strong> the grid quality index to judge a mesh requires knowledge <strong>of</strong> the<br />

range <strong>of</strong> permissible values <strong>of</strong> Q. A calibration is needed which is based on upper<br />

values <strong>of</strong> stretching d x , d y and skewness ω that are considered acceptable for typical<br />

field solvers. The proposed calibration has a conservative character as the upper<br />

bound for index Q is chosen as max(Q x , Q y ). An upper value <strong>of</strong> 20% stretching<br />

1<br />

means a value <strong>of</strong> d x and d y <strong>of</strong> , which when substituted in the expression for Q<br />

11


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 74<br />

0.8<br />

Q x<br />

Normalized EC and Q x<br />

0.6<br />

0.4<br />

0.2<br />

e x xy<br />

e x 2x<br />

e x 2y<br />

0<br />

0 20 40 60 80 100<br />

ω<br />

Figure 3.12: Mesh quality index Q x and its corresponding normalized error coefficients<br />

vs. the skewness angle ω.<br />

yields a value <strong>of</strong> approximately 0.13. Similarly for a maximum skewness angle ω <strong>of</strong><br />

20 ◦ , a value <strong>of</strong> Q ≈ 0.17 is found.<br />

Considering the case <strong>of</strong> a stretched unstructured mesh and applying the same<br />

stretching (center node displacement factor) <strong>of</strong> 20% a value <strong>of</strong> Q ≈ 0.24 is found.<br />

Finally, for the mixed-type element mesh <strong>of</strong> Fig. 3.9(a), setting ψ 2 = tan −1 (2) ± 20 ◦<br />

with ψ 1 = 2 · tan −1 (2) − ψ 2 , an upper bound <strong>of</strong> Q ≈ 0.16 is derived. Therefore, an<br />

upper bound value for Q <strong>of</strong> around 0.20 is a reasonable choice for all types <strong>of</strong> mesh<br />

distortion.<br />

3.9 Application <strong>of</strong> the mesh quality indices<br />

The quality index Q that is based on the first-order error terms appears to be quite<br />

simpler to use compared to the index q that uses both the first and the second<br />

order terms. However, a decision to adopt Q cannot be made until both <strong>of</strong> them<br />

are implemented with general distorted meshes. This section employs three types<br />

<strong>of</strong> grids; structured, unstructured, as well as hybrid. The hybrid grids consist <strong>of</strong><br />

separate quadrilateral and triangular layers, which is quite typical in applications.


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 75<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

Q y<br />

ψ 2<br />

0<br />

30 40 50 60 70 80 90 100<br />

Figure 3.13: Mesh quality index Q y for the mixed-type element mesh interface <strong>of</strong><br />

Fig. 3.9(a) vs. angle ψ 2 .<br />

The geometries involved are channel, airfoil and cylinder. It is important that the<br />

employed grids exhibit (i) both small and large magnitude distortions, as well as (ii)<br />

local and more global ones. The contours <strong>of</strong> the functions Q ≡ max(Q x , Q y ) and<br />

q ≡ max(q x , q y ) are plotted for all cases.<br />

In order to compare the effectiveness—fidelity <strong>of</strong> the indices for the different<br />

cases, the same contour levels are plotted, that is the minimum, maximum and<br />

number <strong>of</strong> levels are the same for all grids for each index. Further, both indices<br />

(Q and q) are shown with the same number <strong>of</strong> contour levels, which is essential for<br />

comparing the two on the same mesh.<br />

3.9.1 Channel grids<br />

The structured channel grid shown in Fig. 3.14(a) is uniform everywhere except<br />

in the middle where a point is displaced. Both indices capture this distortion with<br />

index Q–contours being more focused.<br />

The same displacement is applied to an almost uniform triangular mesh shown


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 76<br />

in Fig. 3.15(a). Similarly here, both indices capture the primary distortion along<br />

with smaller ones with q capturing a bit broader area than Q. This is due to the<br />

second order EC included in the definition <strong>of</strong> index q.<br />

The third mesh is hybrid with a straight line transitioning from the quadrilaterals<br />

to the triangles shown in Fig. 3.16(a). The index Q is more focused on the<br />

interface region than index q does.<br />

3.9.2 Airfoil grids<br />

An O–type structured mesh for the NACA 0012 airfoil is employed next. There is<br />

a sudden expansion <strong>of</strong> the quadrilaterals size in the region close to the surface as<br />

shown in Fig. 3.17. The same observation is made here, namely Q captures the<br />

areas <strong>of</strong> mesh stretching and skewness (leading and trailing edge regions) stronger<br />

compared to index q.<br />

The unstructured mesh around the NACA 0012 shown in Fig. 3.18 is a good<br />

quality mesh with small distortions everywhere. This case was devised to check the<br />

indices’ sensitivity to relatively small (“background”) distortions. It is observed that<br />

both indices show sensitivity to these non–uniformities with q being more sensitive<br />

to them.<br />

Finally, a hybrid mesh around the RAE 2822 airfoil is employed (Fig. 3.19).<br />

This is a viscous mesh with a very good preservation <strong>of</strong> the order <strong>of</strong> accuracy in<br />

the boundary layer region, except at the mesh interfaces, as indicated by index<br />

Q. However, index q shows a large deviation from the ideal computation for the<br />

boundary layer region due to the non alignment <strong>of</strong> the structured mesh in that area<br />

with the axes <strong>of</strong> the system <strong>of</strong> reference.<br />

3.9.3 Cylinder grids<br />

The cylinder mesh has mild distortions away from the body surface. It is an interesting<br />

case to check the performance <strong>of</strong> the quality indices at the farfield. Fig. 3.20


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 77<br />

illustrates the structured mesh which exhibits skewness along lines ”diagonally” <strong>of</strong>f<br />

the surface. The grid is <strong>of</strong> good quality close to the cylinder and this is indicated by<br />

both indices. It is observed that the four “skewness lines” are captured more clearly<br />

by index Q.<br />

The next case involves a triangular grid exhibiting a large stretching area close<br />

to the surface, as well as a mild mesh distortion <strong>of</strong> irregular shape away from the<br />

cylinder (Fig. 3.21). The two indices do recognize both types <strong>of</strong> distortion, and<br />

their contours “follow” the mild distortion pretty closely.<br />

Finally, the hybrid grid <strong>of</strong> Fig. 3.22 exhibits irregularity <strong>of</strong> the interface between<br />

the quadrilaterals and the triangles, as well as quite strong distortions in the<br />

farfield. It is observed that both indices recognize both types <strong>of</strong> strong distortions;<br />

the circular–shaped interface and the “random” ones in the triangles zone.


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 78<br />

(a)<br />

(b)<br />

(c)<br />

Figure 3.14: Locally stretched structured channel mesh: (a) mesh geometry, (b)<br />

index q and (c) index Q.


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 79<br />

(a)<br />

(b)<br />

(c)<br />

Figure 3.15: Locally stretched unstructured channel mesh: (a) mesh geometry, (b)<br />

index q and (c) index Q.


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 80<br />

(a)<br />

(b)<br />

(c)<br />

Figure 3.16: Mixed-type element channel mesh: (a) mesh geometry, (b) index q and<br />

(c) index Q.


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 81<br />

(a)<br />

(b)<br />

(c)<br />

Figure 3.17: O–type structured mesh around a NACA 0012 airfoil: (a) mesh geometry,<br />

(b) index q and (c) index Q.


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 82<br />

(a)<br />

(b)<br />

(c)<br />

Figure 3.18: Unstructured mesh around a NACA 0012 airfoil: (a) mesh geometry,<br />

(b) index q and (c) index Q.


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 83<br />

(a)<br />

(b)<br />

(c)<br />

Figure 3.19: Mixed-type element mesh around a RAE 2822 airfoil: (a) mesh geometry,<br />

(b) index q and (c) index Q.


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 84<br />

(a)<br />

(b)<br />

(c)<br />

Figure 3.20: Structured mesh around a cylinder: (a) mesh geometry, (b) index q<br />

and (c) index Q.


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 85<br />

(a)<br />

(b)<br />

(c)<br />

Figure 3.21: Unstructured mesh around a cylinder: (a) mesh geometry, (b) index q<br />

and (c) index Q.


CHAPTER 3. MESH QUALITY MEASURES FOR THE FV METHOD 86<br />

(a)<br />

(b)<br />

(c)<br />

Figure 3.22: Mixed-type element mesh around a cylinder: (a) mesh geometry, (b)<br />

index q and (c) index Q.


Chapter 4<br />

Discontinuous Galerkin<br />

discretization<br />

The equations governing compressible flow (Eqs. (2.1) and (2.29)) are discretized<br />

with the DG method. The DG method is a FE method which was first introduced<br />

in 1973 by Reed and Hill [122] and further developed to the so called Runge–Kutta<br />

DG (RKDG) method by Cockburn ans Shu in a series <strong>of</strong> papers [28, 40–44]. It uses<br />

discontinuous piecewise polynomials as basis functions and relies on the choice <strong>of</strong><br />

the numerical fluxes, which handle effectively the interactions across the element<br />

interfaces to achieve stable and highly accurate numerical solutions. In the last few<br />

years, the DG method attracted significant interest in the computational mechanics<br />

community, because it <strong>of</strong>fers certain advantages compared to the FV, FD methods<br />

as well as to the classical continuous FE method.<br />

A remarkable advantage <strong>of</strong> this method is that the local support <strong>of</strong> the numerical<br />

solution at the element level, makes it particularly suitable for flow simulations<br />

with high-order <strong>of</strong> accuracy. Moreover, mesh distortion does not affect the quality<br />

<strong>of</strong> the solution considerably as the order <strong>of</strong> the scheme increases [139, 150, 151] to<br />

very high-orders. Furthermore, hp-adaptation is easy to implement permitting high<br />

resolution <strong>of</strong> the flow field at selected flow regions.<br />

Inspired from the DG method other methods with similar features such as<br />

87


CHAPTER 4. DISCONTINUOUS GALERKIN DISCRETIZATION 88<br />

the spectral difference (SD) [98] and the SV [5, 149] methods have been recently<br />

developed. In this work the DG method is used because it is based on the FE<br />

discretization framework and it is straightforward to extend in three dimensions.<br />

The locality <strong>of</strong> the method makes it well suited for parallel implementation. Due<br />

to the large amount <strong>of</strong> computations required for the evaluation <strong>of</strong> the terms appearing<br />

in the semi-discrete form <strong>of</strong> the equations, it is imperative to implement<br />

the algorithms <strong>of</strong> the DG method for parallel computation. The DG method can<br />

easily be implemented on mixed-type element meshes, with the use <strong>of</strong> edge based<br />

data structures.<br />

By way <strong>of</strong> summary the DG method has several potential advantages compared<br />

to other methods including:<br />

ˆ Spectral accuracy on arbitrary meshes.<br />

ˆ Local h-refinement with the use <strong>of</strong> hanging nodes.<br />

ˆ Local p or hp-refinement.<br />

ˆ Boundary conditions are imposed weakly through the numerical flux.<br />

ˆ Local conservation laws allow for different fidelity models on neighboring elements.<br />

ˆ Highly parallelizable.<br />

4.1 Spatial Discontinuous Galerkin discretization<br />

As a FE method the DG discretization relies upon certain definitions, which are<br />

the building blocks <strong>of</strong> all the FE methods. These definitions are provided here in<br />

accordance to Ciarlet [37].<br />

Definition 1 (FE) A finite element is defined as a triplet {Ω e , P e , Σ e }, where :<br />

ˆ Ω e is a subset <strong>of</strong> R d .


CHAPTER 4. DISCONTINUOUS GALERKIN DISCRETIZATION 89<br />

ˆ P e is a vector space <strong>of</strong> functions p e : Ω → R m for some positive integer m.<br />

ˆ Σ e is a set <strong>of</strong> linearly independent linear forms U i , 1 ≤ i ≤ N defined over<br />

the space P e . The set Σ e is P -unisolvent, that is given any real scalar a there<br />

exists a unique function p e ∈ P e , such that U i (p e ) = a.<br />

Definition 2 (FE Mesh) The set T h = {Ω 1 , Ω 2 , . . . , Ω n } over a domain Ω with a<br />

piece wise polynomial boundary defines a tessellation <strong>of</strong> Ω:<br />

Ω = ⋃ e<br />

Ω e .<br />

Each element Ω e has a polynomial order N e .<br />

Definition 3 (FE interpolant) Given a unisolvent finite element {Ω e , P e , Σ e }, let<br />

B = {b 1 , b 2 , . . . , b n } be the set <strong>of</strong> basis that span the space P e . Let v ∈ V where<br />

P e ⊂ V , be a function for which the linear forms U i (p e ) are defined. The local FE<br />

interpolant is defined as:<br />

n∑<br />

U e = c e i b i .<br />

i=1<br />

From the linearity <strong>of</strong> the forms b i , it follows the linearity <strong>of</strong> the interpolation operator<br />

U e .<br />

4.1.1 Discontinuous weak formulation for the Euler equations<br />

A FE mesh T h over an arbitrary domain Ω is considered, where the physical field<br />

is approximated by a function Ũ(x, t), constructed by a locally continuous approximation<br />

(FE interpolant) Ũe(x, t) over each Ω e . For a DG spatial discretization<br />

Ũ(x, t) is considered discontinuous at the inter element boundaries as depicted in<br />

Fig. 4.1. Symbolically, it may be written as:


CHAPTER 4. DISCONTINUOUS GALERKIN DISCRETIZATION 90<br />

Ũ(x, t) = ∑ e<br />

Ũ e (x, t).<br />

Considering a single element Ω e , the approximation Ũe(x, t) substituted in the system<br />

<strong>of</strong> Euler equations (2.1), results in an approximation error, or residual, <strong>of</strong> the<br />

field over the elemental region, equal to:<br />

R e (Ũe(x, t)) = L(Ũe(x, t)), (4.1)<br />

leading to the residual over domain Ω:<br />

R(Ũ(x, t)) = ∑ e<br />

R e (Ũe(x, t)). (4.2)<br />

Ũ h 1<br />

Ũ h 2<br />

Ω 1<br />

Ω 2<br />

Figure 4.1: Local reconstruction <strong>of</strong> the field at two adjacent elements.<br />

As a FE method, the DG discretization, minimizes the approximation error over<br />

the computational domain in the weighted residual sense. That is, forming the<br />

Legendre inner product <strong>of</strong> R(Ũ(x, t)) with an arbitrary test or weighting function<br />

v, and setting it equal to zero:


CHAPTER 4. DISCONTINUOUS GALERKIN DISCRETIZATION 91<br />

∫<br />

Ω<br />

R(Ũ(x, t))vdΩ ≡ (R(Ũ(x, t)), v) = 0,<br />

leads to:<br />

∫<br />

Ω<br />

[<br />

v ∂Ũ<br />

∂t + v∇ · F (Ũ) ]<br />

dΩ = 0 ⇒<br />

∫<br />

Ω<br />

v ∂Ũ<br />

∫<br />

∂t<br />

∫Ω<br />

dΩ − ∇v · F (Ũ)dΩ + ∇ · vF (Ũ)dΩ = 0.<br />

Ω<br />

Using Gauss’ theorem, the last equation results in:<br />

∫<br />

Ω<br />

v ∂Ũ<br />

∮<br />

∂t<br />

∫Ω<br />

dΩ − ∇v · F (Ũ)dΩ + vF (Ũ) · n∂l = 0, (4.3)<br />

∂Ω<br />

where n is the outward normal vector at the domain boundary. Alternatively, Eq.<br />

(4.3) may be written as:<br />

∑<br />

e<br />

[ ∫ ∫<br />

∮<br />

]<br />

v ∂Ũe<br />

Ω e<br />

∂t dΩ e − ∇v · F (Ũe)dΩ e + vF (Ũe) · n e ∂l = 0,<br />

Ω e ∂Ω e<br />

with n e denoting the outward normal vector at the element boundaries. However,<br />

the discontinuous nature <strong>of</strong> the solution Ũ, permits to write the residual for every<br />

element as a separate equation:<br />

Eq.<br />

equations.<br />

∫<br />

Ω e<br />

v ∂Ũe<br />

∂t dΩ e −<br />

∫<br />

∮<br />

∇v · F (Ũe)dΩ e +<br />

Ω e<br />

vF (Ũe) · n e ∂l = 0.<br />

∂Ω e<br />

(4.4)<br />

(4.4) represents the discontinuous weak formulation for the system <strong>of</strong> Euler<br />

The local approximation <strong>of</strong> the field is constructed by a linear combination <strong>of</strong><br />

elemental basis functions b e k (x), according to the previous definitions:


CHAPTER 4. DISCONTINUOUS GALERKIN DISCRETIZATION 92<br />

Ũ e (x, t) = ∑ k<br />

c k e(t)b e k(x), (4.5)<br />

where the coefficient vector c k e(t) denotes the degrees <strong>of</strong> freedom (DOF) <strong>of</strong> the<br />

numerical solution to be advanced in time at every element. Substituting Ũe(x, t)<br />

in Eq. (4.4), yields:<br />

∫<br />

Ω e<br />

vb e k<br />

∂ĉ k ∫<br />

∮<br />

e<br />

∂t dΩ e − ∇v · F (Ũe)dΩ e + vF (Ũe) · n e ∂l = 0, (4.6)<br />

Ω e ∂Ω e<br />

The test function v and the elemental expansion function b e k (x) belong to the same<br />

FE space:<br />

X = {f ∈ L 2 (Ω) : f e ∈ Pe m (Ω e ) ∀Ω e ∈ T h }, (4.7)<br />

where Pe<br />

m (Ω e ) denotes the local space <strong>of</strong> polynomial functions <strong>of</strong> degree at most m<br />

and L 2 (Ω) represents the space <strong>of</strong> functions, which are squared Lebesgue integrable<br />

over Ω e . Then a system <strong>of</strong> 4DOF × 4DOF equations results for the system <strong>of</strong> Euler<br />

equations in two dimensions.<br />

4.1.2 The Numerical flux<br />

An important component for a DG discretization <strong>of</strong> the system <strong>of</strong> Euler equations<br />

is the evaluation <strong>of</strong> the numerical flux. Generally, a numerical flux that satisfies the<br />

following conditions is suitable for a solution algorithm for conservation laws based<br />

on the DG discretization:<br />

ˆ Consistent.<br />

ˆ Conservative.


CHAPTER 4. DISCONTINUOUS GALERKIN DISCRETIZATION 93<br />

The condition <strong>of</strong> consistency requires that the numerical flux H should approximate<br />

the line integral in Eq. (4.21). Typically a requirement <strong>of</strong> Lipschitz continuity is<br />

imposed, so that if a constant D exists the following inequality is satisfied:<br />

)<br />

∣<br />

∣H<br />

(Ũ+ e , Ũ− e<br />

(∣ )<br />

∣∣<br />

− F (U) ∣ ≤ Dmax Ũ + ∣<br />

− U∣ , ∣Ũ− − U∣<br />

. (4.8)<br />

The requirement for conservation in the formulation <strong>of</strong> the numerical flux is mandatory<br />

in order to obtain physically correct solutions for flows with discontinuities.<br />

The numerical fluxes may be chosen based on the exact or approximate Riemann<br />

solvers, developed for the FD and FV methodologies [135]. Qiu et al. in [119]<br />

examined several numerical fluxes starting from the exact Godunov flux up to more<br />

sophisticated fluxes from approximate Riemann solvers, such as the Roe [124] and<br />

the HLLC [136] flux and it was demonstrated that the local Lax-Friedrichs (LLF)<br />

flux is the most efficient in terms <strong>of</strong> CPU time, and that at orders <strong>of</strong> approximation<br />

higher than 3, the behavior <strong>of</strong> the resulting scheme with the LLF flux being almost<br />

the same with other less diffusive, but more computationally expensive numerical<br />

fluxes. This must have been the main reason that most RKDG works in the literature<br />

to be formulated with the LLF flux, as it has been done in the present work.<br />

The LLF flux is:<br />

)<br />

H<br />

(Ũ+ e , Ũ− e = 1 2<br />

[ ] + (fi + f − i )n x + (g + i + g − 1<br />

i )n y −<br />

2 k ( )<br />

U + e − U − e , (4.9)<br />

where the superscripts (+) and (-) denote the exterior and interior elements respectively,<br />

sharing the same edge and k is the spectral radius (maximum eigenvalue) <strong>of</strong><br />

the flux Jacobian given in Eq. (2.24) along the direction n normal to the edge and<br />

is equal to:<br />

k = u · n + c,


CHAPTER 4. DISCONTINUOUS GALERKIN DISCRETIZATION 94<br />

and it is obtained from the flux Jacobian evaluated at the Roe averaged state:<br />

ρu ˜ =<br />

√<br />

ρ+ u + + √ ρ − u −<br />

√<br />

ρ+ + √ ρ − , (4.10)<br />

ρv ˜ =<br />

√<br />

ρ+ v + + √ ρ − v −<br />

√<br />

ρ+ + √ ρ − , (4.11)<br />

˜h =<br />

√<br />

ρ+ h + + √ ρ − h −<br />

√<br />

ρ+ + √ ρ − , (4.12)<br />

˜c = (γ − 1)<br />

√<br />

˜h − 1 2 ( ρu2 ˜ + ρv ˜<br />

2 ), (4.13)<br />

with h = E + p.<br />

The Roe flux [124] has also been implemented. Roe’s flux is:<br />

)<br />

H<br />

(Ũ+ e , Ũ− e = 1 2<br />

[ ] + (fi + f − i )n x + (g + i + g − 1<br />

i )n y −<br />

2 LΛR ( )<br />

U + e − U − e , (4.14)<br />

where L, R are the left and right eigenvectors (Eqs. (2.28) and (2.27)) and Λ is the<br />

matrix <strong>of</strong> eigenvalues given in Eq. (2.25) computed at the Roe average state. The<br />

Harten and Hyman entropy fix is used to avoid the expansion shock through a sonic<br />

point, alternating the eigenvalues in matrix Λ according to:<br />

The quantity ɛ is determined as follows:<br />

λ i = λ2 i + ɛ 2<br />

. (4.15)<br />

2ɛ<br />

ɛ = max ( 0.01, λ i − λ + i , λ− i − λ i<br />

)<br />

. (4.16)<br />

where λ i is the eigenvalue computed using the Roe averaged variables and λ + i , λ− i<br />

are the eigenvalues computed using the elements’ traces. The index i refers to the i<br />

characteristic field <strong>of</strong> the system.


CHAPTER 4. DISCONTINUOUS GALERKIN DISCRETIZATION 95<br />

4.1.3 Basis functions and standard elemental configuration<br />

The choice <strong>of</strong> the basis functions affects the behavior and computational efficiency <strong>of</strong><br />

the DG discretization [83, 150]. Three types <strong>of</strong> basis functions may be constructed,<br />

while satisfying the unisolvency <strong>of</strong> the FE interpolation: modal, nodal and mixed<br />

modal-nodal.<br />

ˆ Modal bases are a set <strong>of</strong> approximating polynomials which have the property<br />

that by increasing the order <strong>of</strong> approximation, all the previous orders are still<br />

included in the approximation set.<br />

For example, the basis defined by the<br />

following set: {1, x, y, xy, x 2 , y 2 } is a modal hierarchical base.<br />

ˆ Nodal bases are a set <strong>of</strong> approximating polynomials, which have the property<br />

<strong>of</strong> being equal to unity at specific points and zero at other points in the approximation<br />

space. Lagrange interpolation polynomials is an example <strong>of</strong> nodal<br />

bases.<br />

ˆ Mixed modal-nodal bases are a combination <strong>of</strong> the above polynomial sets.<br />

When considering <strong>of</strong> applying a p-type refinement to the numerical solution, it<br />

is advantageous to use modal hierarchical expansion bases. Such bases may be<br />

constructed using the orthogonal set <strong>of</strong> Jacobi polynomials [83]. This approach was<br />

followed in the present work. The basis functions are defined over the standard<br />

square elemental configuration, which permits their systematic construction and<br />

implementation <strong>of</strong> any numerical operation involving them.<br />

Application <strong>of</strong> a Gauss type numerical integration, requires the definition <strong>of</strong> a<br />

standard elemental region spanning the domain [−1, 1] × [−1, 1]. In Fig. 4.2, the<br />

transformation <strong>of</strong> the physical quadrilateral to its standard elemental configuration<br />

Ω q st is depicted, and in Fig. 4.3, the transformation <strong>of</strong> the physical triangle to<br />

its standard elemental configuration Ωst t is shown. In this figures the Ω q st region,<br />

where the basis functions are to be constructed and all the numerical operations are<br />

performed is also shown.


CHAPTER 4. DISCONTINUOUS GALERKIN DISCRETIZATION 96<br />

n 2<br />

n 1<br />

C<br />

x = x(n 1 , n 2 )<br />

(-1,1) (1,1)<br />

D<br />

C<br />

D<br />

B<br />

(0,0)<br />

y<br />

q<br />

A<br />

y = y(n 1 , n 2 )<br />

A<br />

(-1,-1)<br />

B<br />

(1,-1)<br />

O<br />

x<br />

p<br />

Ω q e<br />

Ω q st<br />

Figure 4.2: Transformation <strong>of</strong> the physical quadrilateral to the standard quadrilateral<br />

element configuration (η 1 , η 2 ∈ [−1, 1]).<br />

For quadrilateral elements, the basis is constructed using the tensorial product<br />

<strong>of</strong> Legendre polynomials ˆψ p (n), which are subset <strong>of</strong> the Jacobi polynomials.<br />

according to a specific indexing, over Ω q st [83]:<br />

b k (η 1 , η 2 ) = ˆψ p (η 1 ) ˆψ q (η 2 ), −1 ≤ η 1 , η 2 ≤ 1, (4.17)<br />

with:<br />

k = p + q(N + 1), 0 ≤ p, q ≤ N.<br />

If a local approximation <strong>of</strong> order N is applied, then the number <strong>of</strong> bases in X is<br />

equal to (N + 1) 2 .<br />

For triangular elements, the basis is defined over Ωst t and constructed using<br />

the tensorial product <strong>of</strong> Jacobi polynomials over the region Ω q st (Fig. 4.3) [49]:<br />

b k (ξ 1 , ξ 2 ) = Pp 0,0 (η 1 )( 1 − η 2<br />

) p Pq 2p+1,0 (η 2 ), (4.18)<br />

2<br />

with the use <strong>of</strong> collapsed coordinates [83]:


CHAPTER 4. DISCONTINUOUS GALERKIN DISCRETIZATION 97<br />

(-1,1)<br />

C<br />

ξ 2<br />

(0,0)<br />

ξ 1<br />

A<br />

B<br />

(-1,-1)<br />

(1,-1)<br />

x = x(ξ 1 , ξ 2 )<br />

Ω t st<br />

n 1 = 2 1+ξ 1<br />

1−ξ 2<br />

− 1<br />

y = y(ξ 1 , ξ 2 )<br />

n 2 = ξ 2<br />

C<br />

B<br />

x = x(n 1 , n 2 )<br />

(-1,1) (1,1)<br />

D<br />

C<br />

y<br />

A<br />

Ω t e<br />

y = y(n 1 , n 2 )<br />

q<br />

A<br />

(-1,-1)<br />

(0,0)<br />

n 2<br />

n 1<br />

B<br />

(1,-1)<br />

O<br />

x<br />

p<br />

Ω q st<br />

Figure 4.3: Transformation <strong>of</strong> the physical domain triangle to the standard triangular<br />

element (ξ 1 , ξ 2 ∈ [−1, 1] with ξ 1 + ξ 2 ≤ 1) and standard quadrilateral element<br />

configuration (η 1 , η 2 ∈ [−1, 1]).<br />

ξ 1 = (1 + η 1)(1 − η 2 )<br />

2<br />

and according to the following indexing [93]:<br />

− 1, ξ 2 = η 2 ,<br />

k = p + (N + 1)q − q q − 1 , 0 ≤ p, q ≤ N with p + q ≤ N.<br />

2<br />

For a triangular element the number <strong>of</strong> bases for a local approximation <strong>of</strong> order N,<br />

is equal to (N+1)(N+2)<br />

2<br />

.<br />

In Tables 4.1 and 4.2 the analytical form <strong>of</strong> the basis functions, over the standard<br />

quadrilateral and triangular region, for a third order expansion are given, and


CHAPTER 4. DISCONTINUOUS GALERKIN DISCRETIZATION 98<br />

plotted in Figs. 4.4 and 4.5, respectively.<br />

(a) (b) (c)<br />

(d) (e) (f)<br />

(g) (h) (i)<br />

Figure 4.4: Basis functions over the standard square element: (a) (p, q) = (0, 0),<br />

(b) (p, q) = (1, 0), (c) (p, q) = (2, 0), (d) (p, q) = (0, 1), (e) (p, q) = (1, 1), (f)<br />

(p, q) = (2, 1), (g) (p, q) = (0, 2), (h) (p, q) = (1, 2), (i) (p, q) = (2, 2).


CHAPTER 4. DISCONTINUOUS GALERKIN DISCRETIZATION 99<br />

p, q b k p, q b k p, q b k<br />

(0, 0) 1 (1, 0) η 1 (2, 0)<br />

3n 2 1 −1<br />

2<br />

(0, 1) η 2 (1, 1) η 1 η 2 (2, 1)<br />

3n 2 1 +1<br />

2<br />

η 2<br />

(0, 2)<br />

3n 2 2 −1<br />

2<br />

(1, 2) η 1<br />

3n 2 2 +1<br />

2<br />

(2, 2)<br />

(3n 2 1 +1) (3n 2 2 +1)<br />

2 2<br />

Table 4.1: Basis functions for a third-order approximation over the standard quadrilateral<br />

region.<br />

p, q b k p, q b k p, q b k<br />

1−η<br />

(0, 0) 1 (1, 0) η 2<br />

3<br />

1 (2, 0) (η 2 2 1 − 1) 2 + 3η 1 − 2( 1−η 2<br />

) 2<br />

2<br />

(0, 1)<br />

3η 2 +1<br />

1−η<br />

(1, 1) η 2 5η 2 +3<br />

5<br />

2 1 (0, 2) (η 2 2 2 2 − 1) 2 + 6η 2 − 3<br />

Table 4.2: Basis functions for a third-order approximation over the standard triangular<br />

region.<br />

(a) (b) (c)<br />

(d) (e) (f)<br />

Figure 4.5: Basis functions over the standard triangular region: (a) (p, q) = (0, 0),<br />

(b) (p, q) = (1, 0), (c) (p, q) = (2, 0), (d) (p, q) = (0, 1), (e) (p, q) = (1, 1), (f)<br />

(p, q) = (0, 2).


CHAPTER 4. DISCONTINUOUS GALERKIN DISCRETIZATION 100<br />

4.1.4 Elemental operations<br />

In order to handle arbitrary geometries, it is necessary to numerically evaluate all<br />

the integrals appearing in Eq. (4.6). These are the mass matrix M kj , the volume<br />

integral V k and the surface (line) integral S k :<br />

M kj =<br />

∫<br />

b e kb e jdΩ e ,<br />

Ω e<br />

(4.19)<br />

∫<br />

V k = ∇b e k · F (Ũe)dΩ e ,<br />

Ω e<br />

(4.20)<br />

S k =<br />

∮<br />

b e kF (Ũe) · n e ∂l,<br />

∂Ω e<br />

(4.21)<br />

where in place <strong>of</strong> the test function v the elemental basis function b e j has been substituted.<br />

In the present work, Gaussian quadrature rules are used for numerical<br />

integration.<br />

For quadrilateral elements, the integrals given in Eqs. (4.19) and (4.20) are<br />

then computed by the weighted summations (η = η 1,i , η 2,m ):<br />

∫ 1 ∫ 1<br />

M kj = b k b j |J|dη 1 dη 2 =<br />

−1 −1<br />

Q<br />

∑ 1 Q<br />

∑ 2<br />

w i w m b k (η)b j (η)|J(η)|,<br />

i=1 m=1<br />

(4.22)<br />

∫ 1 ∫ 1<br />

V k = J −1 · ∇b k · F (Ũ e )|J|dη 1 dη 2 =<br />

−1 −1<br />

Q<br />

∑ 1 Q<br />

∑ 2<br />

w i w m J −1 (n) · ∇b k (η) · F (Ũ e (η)|J(η)|,<br />

i=1 m=1<br />

(4.23)


CHAPTER 4. DISCONTINUOUS GALERKIN DISCRETIZATION 101<br />

and for triangular elements, the same integrals are computed by the following<br />

weighted summations:<br />

∫ 1<br />

M kj =<br />

∫ −ξ2<br />

−1 −1<br />

∫ 1 ∫ 1<br />

−1 −1<br />

Q<br />

∑ 1 Q<br />

∑ 2<br />

i=1 m=1<br />

b k b j |J|dξ 1 dξ 2 =<br />

b k b j |J|( 1 − η 2<br />

)dη 1 dη 2 =<br />

2<br />

( 1 − η2,m<br />

w i w m b k (η)b j (η)|J(η)|<br />

2<br />

)<br />

,<br />

(4.24)<br />

∫ 1<br />

V k =<br />

∫ −ξ2<br />

−1 −1<br />

∫ 1 ∫ 1<br />

−1 −1<br />

Q<br />

∑ 1 Q 2<br />

i=1 m=1<br />

J −1 · ∇b k · F (Ũ e )|J|dξ 1 dξ 2 =<br />

J −1 · ∇b k · F (Ũ e )|J|( 1 − η 2<br />

)dη 1 dη 2<br />

2<br />

∑<br />

( 1 −<br />

w i w m J −1 η2,m<br />

(n) · ∇b k (η) · F (Ũ e (η))|J(n)|<br />

2<br />

)<br />

,<br />

(4.25)<br />

where |J| is the Jacobian <strong>of</strong> the transformation, J −1 is the inverse <strong>of</strong> the Jacobian<br />

matrix <strong>of</strong> the transformation from the physical to the computational space, w i , w m<br />

are the quadrature weights and Q 1 and Q 2 is the number <strong>of</strong> quadrature points<br />

along the η 1 and η 2 direction respectively. Matrix M is called the mass matrix and<br />

possesses a diagonal structure only for straight sided triangles and for straight sided<br />

quadrilaterals with parallel edges.<br />

The line integral in Eq. (4.21) is computed by the following weighted summation,<br />

both for triangles and quadrilaterals:


CHAPTER 4. DISCONTINUOUS GALERKIN DISCRETIZATION 102<br />

S k =<br />

Q m<br />

∑<br />

m=1<br />

∫ 1<br />

−1<br />

b k F (Ũe) · n e |l e |dn =<br />

w m b k H(Ũ+ e , Ũ− e ) · n c e|l e |,<br />

(4.26)<br />

where Q m is the number <strong>of</strong> quadrature points, |l e | is the edge length, H the numerical<br />

flux at the element boundary, Ũ+ e and Ũ− e are the element traces on an edge shared<br />

by two elements and n c e is the normalized outward vector <strong>of</strong> the local element edge<br />

at the computational space.<br />

Clearly, in our implementation all numerical computations <strong>of</strong> the DG method<br />

are carried out in the canonical computational domain <strong>of</strong> the square element. This<br />

choice gives a significant advantage for use <strong>of</strong> mixed-type element meshes and the<br />

novel implementation <strong>of</strong> the limiters for arbitrary elements that will be shown later.<br />

4.1.5 Selection <strong>of</strong> Gauss type numerical integration<br />

Gauss numerical integration is classified according to the way the end points <strong>of</strong> the<br />

domain <strong>of</strong> integration are handled [83]. The classification is given in Table 4.3, with<br />

the corresponding order <strong>of</strong> accuracy using Q n quadrature points.<br />

Type End points Order <strong>of</strong> accuracy<br />

Gauss-Legendre (GL) no end points 2Q n − 1<br />

Gauss-Radau (GR) only one end point 2Q n − 2<br />

Gauss-Lobatto-Legendre (GLL) both end points 2Q n − 3<br />

Table 4.3: Types <strong>of</strong> Gauss numerical integration.


CHAPTER 4. DISCONTINUOUS GALERKIN DISCRETIZATION 103<br />

The GL and GLL type have been implemented for both the surface and the<br />

line integrations in the present work. Moreover, in Eqs. (4.22), (4.24) and (4.26),<br />

it is seen that the product <strong>of</strong> two polynomials <strong>of</strong> order N is integrated. Thus, the<br />

same number <strong>of</strong> quadrature points is used for volume and surface integrations and<br />

equal to:<br />

Q 1 = Q 2 = Q m = N + 1, (4.27)<br />

for a GL integration, and:<br />

Q 1 = Q 2 = Q m = N + 2, (4.28)<br />

for a GLL integration.<br />

4.1.6 Mapping between physical and computational space<br />

The computation <strong>of</strong> the integrals in Eqs. (4.22) to (4.25) requires the evaluation <strong>of</strong><br />

the Jacobian matrix <strong>of</strong> the transformation at every quadrature point. This necessitates<br />

the construction <strong>of</strong> a mapping from the physical to the computational space.<br />

Usually, according to the polynomial degree N B <strong>of</strong> the physical space representation,<br />

curved elements are classified in sub-, iso- and super-parametric elements, by<br />

comparing the order N B with that <strong>of</strong> the order N <strong>of</strong> the discretization scheme. In<br />

detail:<br />

N B < N : sub-parametric element<br />

N B = N : iso-parametric element<br />

N B > N : super-parametric element<br />

For straight sided quadrilateral elements, the mapping from the physical configuration<br />

<strong>of</strong> the element Ω e to the standard elemental region (see Fig. 4.2), is given<br />

by the following relations:


CHAPTER 4. DISCONTINUOUS GALERKIN DISCRETIZATION 104<br />

(1 − η 1 ) (1 − η 2 )<br />

x =x A<br />

2 2<br />

(1 − η 1 ) (1 + η 2 )<br />

+ x D<br />

2 2<br />

(1 + η 1 ) (1 − η 2 )<br />

+ x B<br />

2 2<br />

+ x C<br />

(1 + η 1 )<br />

2<br />

(1 + η 2 )<br />

,<br />

2<br />

(4.29)<br />

(1 − η 1 ) (1 − η 2 )<br />

y =y A<br />

2 2<br />

(1 − η 1 ) (1 + η 2 )<br />

+ y D<br />

2 2<br />

(1 + η 1 ) (1 − η 2 )<br />

+ y B<br />

2 2<br />

+ y C<br />

(1 + η 1 )<br />

2<br />

(1 + η 2 )<br />

,<br />

2<br />

(4.30)<br />

and for straight sided triangular elements (see Fig. 4.3) by:<br />

(1 − η 1 ) (1 − η 2 )<br />

x = x A<br />

2 2<br />

(1 + η 1 ) (1 − η 2 )<br />

+ x B<br />

2 2<br />

1 + η 2<br />

+ x C , (4.31)<br />

2<br />

(1 − η 1 ) (1 − η 2 )<br />

y = y A<br />

2 2<br />

(1 + η 1 ) (1 − η 2 )<br />

+ y B<br />

2 2<br />

The Jacobian matrix <strong>of</strong> the transformation is defined as:<br />

1 + η 2<br />

+ y C . (4.32)<br />

2<br />

⎡<br />

⎢<br />

J = ⎣<br />

∂x<br />

∂η 1<br />

∂y<br />

∂η 2<br />

⎤<br />

∂x<br />

∂η 2<br />

⎥<br />

∂y<br />

∂η 2<br />

⎦ , (4.33)<br />

with its inverse equal to:<br />

J −1 = 1<br />

|J|<br />

⎡<br />

⎢<br />

⎣<br />

∂y<br />

∂η 2<br />

− ∂y<br />

∂η 1<br />

− ∂x<br />

∂η 2<br />

∂x<br />

∂η 1<br />

⎤<br />

⎥<br />

⎦ . (4.34)<br />

In the present work, in order to reduce the amount <strong>of</strong> operations required for<br />

the evaluation <strong>of</strong> the surface integrals, the product <strong>of</strong> the elements <strong>of</strong> the inverse


CHAPTER 4. DISCONTINUOUS GALERKIN DISCRETIZATION 105<br />

Jacobian <strong>of</strong> the transformation with the gradient <strong>of</strong> the weighting functions and<br />

the quadrature weights at each quadrature point, is computed and stored in a preprocessing<br />

step.<br />

4.2 DG discretization for the Navier-Stokes Equations<br />

Following the same procedure for the derivation <strong>of</strong> the weak form for the system<br />

<strong>of</strong> the Euler equations, Eq. (4.6), in the DG discretization, the weak form for the<br />

system <strong>of</strong> Navier-Stokes equations is the following:<br />

∫<br />

vb e ∂ĉ k ∫<br />

∮<br />

e<br />

k<br />

Ω e<br />

∂t dΩ e − ∇v · F (Ũe)dΩ e + vF (Ũe) · n e ∂l =<br />

Ω e ∂Ω<br />

∫<br />

∮<br />

e<br />

− ∇v · F v (Ũe, ∇Ũe)dΩ e + vF v (Ũe, ∇Ũe) · n e ∂l.<br />

Ω e ∂Ω e<br />

(4.35)<br />

From the previous system it is observed that the inclusion <strong>of</strong> the viscous and thermal<br />

conduction effects has lead to a system including the solution gradient. This<br />

complicates the solution <strong>of</strong> the Navier-Stokes system in the DG framework, since,<br />

the solution gradient has to be computed and appropriate numerical fluxes for the<br />

viscous part must be chosen to provide a stable and accurate discretization scheme.<br />

4.2.1 Computation <strong>of</strong> the viscous terms in the DG method<br />

The computation <strong>of</strong> the viscous terms in the DG method requires a special attention<br />

as has been demonstrated in [155]. Specifically, considering the one dimensional heat<br />

equation:<br />

u t − u xx = 0, (4.36)


CHAPTER 4. DISCONTINUOUS GALERKIN DISCRETIZATION 106<br />

with periodic boundary conditions and following the same procedure for the derivation<br />

<strong>of</strong> the weak form in the DG method as for the purely hyperbolic system <strong>of</strong> the<br />

Euler equations, yields:<br />

∫<br />

E<br />

v du h<br />

dt<br />

∮∂E<br />

dE + v du h<br />

dx<br />

∫E<br />

de − dv<br />

dx<br />

du h<br />

dE = 0. (4.37)<br />

dx<br />

where u h denotes the solution expansion <strong>of</strong> u. It is obvious that someone now has<br />

to deal with the derivative du h<br />

in the computation <strong>of</strong> the integrals appearing in<br />

dx<br />

Eq. (4.37). Due to the absence <strong>of</strong> an upwind mechanism in the heat equation, the<br />

natural choice is to define the numerical flux for u x as follows:<br />

(<br />

û x = 1 (duh ) +<br />

+<br />

2 dx<br />

( ) ) − duh<br />

. (4.38)<br />

dx<br />

If a P 0 expansion <strong>of</strong> the solution was chosen then this would lead to the analytical<br />

value <strong>of</strong> u h x to be equal to zero and hence the discretization would be innately<br />

inconsistent. Furthermore, in [155] it was demonstrated that this holds for orders<br />

<strong>of</strong> approximation greater than 0.<br />

Several schemes for the discretization <strong>of</strong> the viscous terms have been proposed<br />

in the literature, such as the schemes developed by Bassi and Rebay [17, 19, 20],<br />

the Local DG method (LDG) <strong>of</strong> Cockburn and Shu [43], which is a generalization<br />

<strong>of</strong> the ideas <strong>of</strong> Bassi and Rebay, and the method developed by Baumann and Oden<br />

(BO) [22, 23] and [107]. In the BO DG method, extra terms were added in the<br />

numerical flux that accounted for the jump in the solution between the elements,<br />

but for a P 0 solution expansion the scheme is inconsistent. Moreover, for a P N order<br />

solution expansion, the order <strong>of</strong> accuracy <strong>of</strong> the BO scheme is N for even N, that<br />

is sub-optimal, and N + 1 for odd N. Furthermore, in [155] it was demonstrated<br />

that the BO method has larger errors than the LDG method. For those reasons,<br />

in this work the LDG method is used for discretizing the system <strong>of</strong> Navier-Stokes<br />

equations in a consistent, stable and accurate way.


CHAPTER 4. DISCONTINUOUS GALERKIN DISCRETIZATION 107<br />

4.2.2 The LDG method<br />

The basic idea <strong>of</strong> the LDG method is to perform an order reduction <strong>of</strong> the initial<br />

system into a degenerate first order system and then discretize it within the DG<br />

framework to compute the solution gradient. For doing so, the following auxiliary<br />

variable Θ e at each element is introduced:<br />

⎡ ⎤<br />

∇ρ e<br />

Θ e = ∇Ũe =<br />

∇u e<br />

⎢<br />

⎣∇v ⎥<br />

e ⎦ , (4.39)<br />

∇T e<br />

which in essence is the solution gradient at each element. Multiplying the above<br />

equation by the same weighting function v appearing in Eq. (4.35) and integrating<br />

over the elemental space Ω e , obtain:<br />

∫<br />

∫<br />

vΘ e dΩ e −<br />

Ω e<br />

Ω e<br />

v<br />

⎡ ⎤<br />

∇ρ e<br />

∇u e<br />

⎢<br />

⎣∇v ⎥<br />

e ⎦ dΩ e = 0,<br />

∇T e<br />

which after application <strong>of</strong> Gauss’ theorem, leads to:<br />

⎡ ⎤<br />

ρ e<br />

∫<br />

∫<br />

∮<br />

vΘ e dΩ e + ∇v<br />

u e<br />

⎢<br />

Ω e Ω e ⎣v ⎥<br />

e ⎦ dΩ e −<br />

T e<br />

∂Ω e<br />

v<br />

⎡ ⎤<br />

ρ e<br />

u e<br />

⎢<br />

⎣v ⎥<br />

e ⎦ n e∂l = 0. (4.40)<br />

T e<br />

The evaluation <strong>of</strong> the line integral in Eq. (4.40) along with the numerical viscous<br />

flux computation has a pivotal role in the accuracy and stability <strong>of</strong> the resulting DG<br />

scheme.<br />

Arnold et. al in [6] performed an analysis <strong>of</strong> the resulting schemes depending<br />

on the choice <strong>of</strong> the line integral evaluation. Furthermore, in [6] it was proven


CHAPTER 4. DISCONTINUOUS GALERKIN DISCRETIZATION 108<br />

that two main requirements have to be met by the numerical fluxes in order to<br />

obtain a stable scheme, which furthermore achieves an optimal rate <strong>of</strong> convergence.<br />

These requirements are that the fluxes have to be consistent and conservative. The<br />

examined formulations are given in Table 4.4, where the following notation <strong>of</strong> jump<br />

[] and averaging {} operators for scalar and vector fields is used.<br />

No. Method û ˆΘ<br />

1 Bassi-Rebay (BR1) {u h } {Θ}<br />

2 Brezzi et. al 1 {u h } {Θ} − a r [u h ]<br />

3 LDG {u h } − β · [u h ] {Θ} + β[Θ] − a j [u h ]<br />

4 IP {u h } {∇ h u h } − a j [u h ]<br />

5 Bassi et al. (BR2) {u h } {∇ h u h } − a r [u h ]<br />

6 Baumann-Oden {u h } + n κ · [u h ] {∇ h u h }<br />

7 NIPG {u h } + n κ · [u h ] {∇ h u h } − a j [u h ]<br />

8 Babuska-Zlamal (u h | K )∂K −a j [u h ]<br />

9 Brezzi et. al 2 (u h | K )∂K −a j [u h ]<br />

Table 4.4: DG methods and their interiors fluxes taken for Arnold et. al [6].<br />

ˆ for a scalar variable s:<br />

[s] = s + n + + s − n − = n(s + − s − ),<br />

{s} = 1 2<br />

(<br />

s + + s −) ,


CHAPTER 4. DISCONTINUOUS GALERKIN DISCRETIZATION 109<br />

ˆ for a vector variable g:<br />

[g] = g + · n + + g − · n − = n(g + − g − ),<br />

{g} = 1 2<br />

(<br />

g + + g −) .<br />

The functional operators a r and a j in Table 4.4 are penalty terms which their appearance<br />

in the gradient variable Θ is <strong>of</strong> significant importance for obtaining a stable<br />

scheme. The methods which do not contain any <strong>of</strong> these terms are only weakly stable<br />

and it is obvious that the current chosen LDG method may be or may not be<br />

weakly stable. All the methods with penalty terms may be interpreted as interior<br />

penalty (IP) methods. The penalty method generally imposes a penalty term in<br />

the formulation in order to enforce the solution and must automatically satisfy the<br />

boundary conditions. This is achieved by the presence <strong>of</strong> the penalty term which<br />

adds more artificial dissipation to the jump at the boundary so that to smear out<br />

the discontinuity that appears there.<br />

In the present work, the utilization <strong>of</strong> the LDG method permits, as it easily<br />

verified from Table 4.4, the application <strong>of</strong> the BR1 scheme which in the literature<br />

is known as the first Bassi-Rebay scheme. However, Cockburn and Shu in [43]<br />

demonstrated some deficiencies <strong>of</strong> the BR1 scheme:<br />

ˆ Convergence order <strong>of</strong> N only for odd weighting function.<br />

ˆ Spread stencil.<br />

Both problems may be remedied by using an upwind like fashion on the choice <strong>of</strong><br />

fluxes in the gradient computation and in the viscous flux evaluation. In the present<br />

work this is achieved by the following use <strong>of</strong> the parameters β = ±0.5 and a j = 0.


CHAPTER 4. DISCONTINUOUS GALERKIN DISCRETIZATION 110<br />

Recently in the literature appeared the so called hybridizable DG (HDG) [38,<br />

39, 104] methods, which aim at alleviating the high computational cost <strong>of</strong> the DG<br />

discretization for the viscous terms. In HDG methods the final system is produced<br />

in terms <strong>of</strong> the DOF <strong>of</strong> the approximate traces <strong>of</strong> the field variables, that is the<br />

solution approximation at the inter element boundaries. Since the approximate<br />

traces are defined at the edges/faces only and are single valued, the HDG methods<br />

include significantly fewer coupled unknowns than other DG methods for higher<br />

order terms.<br />

The HDG methods are devised using a hybridization procedure <strong>of</strong> the DG<br />

methodology to discretize a system <strong>of</strong> equations. This procedure starts by introducing<br />

first the approximate traces and then define a total numerical flux in terms<br />

<strong>of</strong> them. A continuity requirement is weakly imposed for the normal component <strong>of</strong><br />

the numerical flux and this leads to a large nonlinear system <strong>of</strong> equations for the<br />

approximate variables (including their spatial gradient) and the numerical fluxes.<br />

After linearization <strong>of</strong> the final system the local variables and their spatial gradient<br />

may be condensed in an element by element fashion and obtain a reduced global<br />

system involving only the approximate traces as the unknowns.<br />

4.3 The treatment <strong>of</strong> initial and boundary conditions<br />

According to the flow conditions that have to be met at specific regions <strong>of</strong> the<br />

computational domain, appropriate boundary conditions (BC) have to be applied<br />

to the system <strong>of</strong> the governing equations. In the present work, the implementation<br />

<strong>of</strong> Neumann and Dirichlet BC is performed in a weak manner. This follows from the<br />

flux formulation <strong>of</strong> the present DG discretization leading to a natural weak inclusion<br />

<strong>of</strong> the BC through the numerical fluxes. It is noted that the weak enforcement <strong>of</strong><br />

Dirichlet BC leads to solutions that still ”feel” the influence <strong>of</strong> the boundary through<br />

the numerical fluxes, but in a manner that is consistent with the accuracy <strong>of</strong> the<br />

interior solution, thus, leading to improved solutions away from the boundaries [9,


CHAPTER 4. DISCONTINUOUS GALERKIN DISCRETIZATION 111<br />

45]. Also, the application <strong>of</strong> Neumann BC comes naturally in the DG discretization<br />

again through the flux formulation <strong>of</strong> the scheme.<br />

The technique for weak enforcement <strong>of</strong> BC requires the use <strong>of</strong> ghost elements,<br />

where a solution state U g is determined such that the appropriate boundary values<br />

at every quadrature point over a boundary edge are set. Then, the underlying<br />

Riemann problem is solved according to the main discretization scheme.<br />

It is noted, however, that a polyline representation <strong>of</strong> the wall geometry is<br />

not appropriate for the DG discretization [16, 18, 92, 93], as the DG method is<br />

highly sensitive to the accuracy <strong>of</strong> the boundary representation. This necessitates<br />

either use <strong>of</strong> curved elements or refinement <strong>of</strong> the mesh at the regions where the<br />

wall curvature is high. In the present work, the last approach is followed as only<br />

straight sided elements are used.<br />

The application <strong>of</strong> the initial conditions (IC) in the DG method may be performed<br />

in two ways that will be explained in the current chapter. These are the<br />

collocation and the discrete Galerkin projection. Both may be applied to solution<br />

expansions <strong>of</strong> nodal or modal hierarchical bases, but the first is more suitable for<br />

solution expansions with nodal bases which approximate the solution at specific<br />

points in the computational element, while the second is more suitable for modal<br />

hierarchical bases.<br />

4.4 Wall boundary conditions<br />

For a flow simulation over an arbitrary geometry, the BCs that have to be satisfied<br />

at the boundary segments defining the geometry depend on the physical modeling<br />

being performed for the flow field. Specifically, for inviscid flow numerical solutions<br />

it is required the flow to be tangent at the geometry, leading to the employment <strong>of</strong><br />

the so called slip-wall BC, which is a Dirichlet type condition and simply expresses<br />

that the flow cannot penetrate through the segments defining the geometry. For the<br />

DG discretization this means that at the quadrature points on the edges defining the<br />

wall geometry the velocity field must be enforced to be tangent. This is expressed


CHAPTER 4. DISCONTINUOUS GALERKIN DISCRETIZATION 112<br />

by the following Dirichlet BC:<br />

v · n = 0, (4.41)<br />

The weak enforcement <strong>of</strong> the BC condition <strong>of</strong> tangency in the velocity field is implemented<br />

at every quadrature point at the wall edges by the solution <strong>of</strong> a Riemann<br />

problem. Specifically, a ghost state is created at the opposite side <strong>of</strong> the computational<br />

domain at a wall edge, where the density, pressure and the magnitude <strong>of</strong> the<br />

velocity field are extrapolated from the interior solution. This corresponds to the<br />

following Neumann BC for the pressure at the wall:<br />

∂p<br />

∂n = 0.<br />

Then the velocity components at the ghost state are determined so that the averaged<br />

velocity field at the quadrature points to be tangent to the edge. Analytically, the<br />

ghost state is the following:<br />

ρ g = ρ,<br />

u g = u t cos (a) + u n sin (a) ,<br />

v g = −u n cos (a) + u t sin (a) ,<br />

(4.42)<br />

P g = P,<br />

where u t , u n are the velocity components tangent and normal to the edge and a is<br />

the angle formed by the wall segment with the x coordinate axis.<br />

For viscous flow computations, the flow must adhere to the geometry resulting<br />

to the application <strong>of</strong> another Dirichlet condition, namely the no-slip wall condition.<br />

The weak enforcement <strong>of</strong> the no-slip wall condition is performed by constructing the


CHAPTER 4. DISCONTINUOUS GALERKIN DISCRETIZATION 113<br />

following ghost state:<br />

ρ g = ρ,<br />

u g = −u t cos (a) + u n sin (a) ,<br />

v g = −u n cos (a) − u t sin (a) ,<br />

(4.43)<br />

P g = P,<br />

and the underlying Riemann problem is solved. Furthermore, for viscous computations<br />

additional considerations have to be taken regarding the heat flux in the<br />

flow field at the boundaries <strong>of</strong> the computational domain. This coordinates the use<br />

either adiabatic, isothermal etc. conditions that are enforced to the Navier-Stokes<br />

system through Neumann or Dirichlet BC. For instance, an adiabatic wall requires<br />

that there is no heat flux normal to the wall and this is expressed as follows:<br />

dT<br />

dn<br />

= 0, (4.44)<br />

and implemented by requiring at the quadrature points on the wall edges the temperature<br />

gradient normal to the wall to be zero. For isothermal walls or for flows<br />

with specific temperature distribution on the wall boundary, the temperature field<br />

must be set at every quadrature point to the prescribed values.<br />

4.5 Far field boundary conditions<br />

At the regions <strong>of</strong> the computational domain where the inlet and outlet <strong>of</strong> the flow<br />

is discretized, BC depending on the local characteristic structure <strong>of</strong> the flow field<br />

must be applied [64, 113, 140]. In the current DG discretization, their application,<br />

also follows that <strong>of</strong> the weak enforcement, and they are based on the assumptions<br />

<strong>of</strong> a locally one dimensional isentropic flow. Also, the application <strong>of</strong> the far field BC


CHAPTER 4. DISCONTINUOUS GALERKIN DISCRETIZATION 114<br />

is performed on the primitive variables [ρ, u, v, p] T .<br />

Considering the system <strong>of</strong> Euler equations and their eigenvalues given in Eq.<br />

(2.25), it is observed that for a supersonic flow all the characteristic waves are<br />

traveling in one direction. This dictates the conditions at the inlet to be imposed<br />

by the conditions <strong>of</strong> the free stream flow and at the outlet all the conditions to be<br />

extrapolated from the interior <strong>of</strong> the domain. However, for the case <strong>of</strong> subsonic flow,<br />

the inlet and outlet regions are crossed by waves coming both from the interior and<br />

the exterior <strong>of</strong> the domain. This situation necessitates modeling <strong>of</strong> the waves crossing<br />

the far field boundaries from the exterior, in order to determine the appropriate BC<br />

at subsonic inlets and outlets. The Riemann invariants [64] are used for determining<br />

the far field BC according to the previously made assumptions:<br />

J + = u + + 2<br />

γ − 1 c+ ,<br />

J − = u − − 2<br />

γ − 1 c− ,<br />

(4.45)<br />

where the + and - superscript refer to the appropriate side <strong>of</strong> the inlet or outlet<br />

region. The primitive, and thus the conservative variables, are then calculated according<br />

to the following algorithm:<br />

1: u b = J + +J −<br />

2<br />

2: c b = g−1<br />

4<br />

(J + − J − )<br />

3: p b = ρc2<br />

γ<br />

4: S b = p b<br />

ρ γ<br />

( ) 1<br />

c 3 γ<br />

5: ρ b = b<br />

γS


CHAPTER 4. DISCONTINUOUS GALERKIN DISCRETIZATION 115<br />

4.6 Application <strong>of</strong> Initial Conditions<br />

The use <strong>of</strong> a modal bases set requires determination <strong>of</strong> the initial values <strong>of</strong> the<br />

DOF according to the IC over the computational domain. Two methods may be<br />

used for the initialization <strong>of</strong> the DOF: collocation projection and discrete Galerkin<br />

projection.<br />

Considering a two dimensional function Ũ(x, t) which does not lie within the<br />

polynomial space <strong>of</strong> the expansion basis, then there will be an error R e (Ũ) in the<br />

approximation <strong>of</strong> the function over an arbitrary element where a solution expansion<br />

Ũ e (x, t) is employed:<br />

R e (Ũ) = Ũ(x, t) − Ũe(x, t). (4.46)<br />

Following the method <strong>of</strong> weighted residuals the Legendre inner product <strong>of</strong> Eq.<br />

(4.46) with an arbitrary function v is formed:<br />

∫<br />

Ω<br />

∫<br />

(Ũ(x, t) − Ũ e (x, t))<br />

vdΩ = R e (U)vdΩ, (4.47)<br />

Ω<br />

and setting the right hand side equal to zero, the following system is obtained for<br />

the determination <strong>of</strong> the expansion coefficients:<br />

∫<br />

Ω<br />

If the function v in Eq.<br />

∫<br />

Ũ e (x, t)vdΩ =<br />

Ω<br />

Ũ(x, t)vdΩ. (4.48)<br />

(4.47) is set equal to the Dirac delta function at<br />

the quadrature points, it is then implied that ∫ R(u)dΩ = 0 and the expansion<br />

Ω<br />

coefficients are obtained by:<br />

Û k e = B −1 · U k e, (4.49)<br />

where the matrix B i,j = b i (η j ). This approach corresponds to the collocation projection.<br />

If the function v is set equal to the basis function b i then the discrete


CHAPTER 4. DISCONTINUOUS GALERKIN DISCRETIZATION 116<br />

Galerkin projection is obtained with B i,j = ∫ b Ω i(η)b j (η)dΩ. In the present work<br />

the Galerkin projection has been used for applying the IC due to the use <strong>of</strong> a modal<br />

bases set.


Chapter 5<br />

Time discretization<br />

For time advancement <strong>of</strong> the compressible flow equations either the method <strong>of</strong> lines<br />

or the space-time DG method could be used. For space-time DG discretizations<br />

the solution is not only discretized in space but in time as well. This leads to<br />

a discretization scheme where the DOFs are scalar but the basis functions <strong>of</strong> the<br />

solution expansion depend both in space and time. For more details in the spacetime<br />

DG discretization refer to [86, 143].<br />

For the method <strong>of</strong> lines, the space discretization is performed first and the<br />

time is considered continuous. Then, a set <strong>of</strong> ODEs <strong>of</strong> size equal to the number <strong>of</strong><br />

DOFs in Eq. (4.6) for the Euler equations and in Eqs. (4.35) and (4.40) for the<br />

Navier-Stokes equations in the current implementation is obtained, which can be<br />

advanced in time either with explicit or implicit methods for ODEs.<br />

Use <strong>of</strong> explicit time integration schemes has been widespread in DG discretizations.<br />

However, explicit schemes suffer from severe stability limitations. The allowable<br />

time step due to CFL-stability limits is very small and few cases may be solved<br />

at a reasonable computing time. On the other hand, the application <strong>of</strong> implicit time<br />

integration does not have limitations imposed by CFL-stability limits, at least for<br />

linear conservation laws, and thus, can allow larger time steps, which are <strong>of</strong>ten determined<br />

only by accuracy considerations. Moreover, implicit methods are suitable<br />

for stiff problems.<br />

117


CHAPTER 5. TIME DISCRETIZATION 118<br />

5.1 Runge-Kutta methods<br />

Runge-Kutta (RK) methods belong to a class <strong>of</strong> multistage single step methods.<br />

They apply numerical integration rules over a specified set <strong>of</strong> stages, where progressively<br />

more accurate approximation <strong>of</strong> the solution’s slope is performed. They are<br />

classified as explicit and implicit, with the solution at every stage being well defined<br />

for small enough step size. Although, their implementation is expensive in terms<br />

<strong>of</strong> storage and processing, they are widely applied in computational mechanics, due<br />

to their exceptional properties in terms <strong>of</strong> stability accompanied with high-order<br />

accuracy.<br />

The generic form <strong>of</strong> an initial value problem for a scalar equation is the following:<br />

dy<br />

=f (t, y (t)) , t ∈ [a, b] ,<br />

dt<br />

y (a) = y 0 ,<br />

(5.1)<br />

and the numerical solution sought is denoted as y : [a, b] → R. Next, defining the<br />

following constants: q ∈ N, τ i ∈ R, a ij ∈ R, b i ∈ R with i, j = 1, . . . , q, and usually<br />

0 ≤ τ i ≤ 1 for and arbitrary function ψ (s) : R → R the following integrals:<br />

∫ ti<br />

0<br />

ψ (s) ds,<br />

∫ 1<br />

0<br />

ψ (s) ds,<br />

are approximated by the following summations, respectively:<br />

q∑<br />

a ij ψ (s) ,<br />

j=1<br />

q∑<br />

b i ψ (τ i ) , i = 1, . . . , q.<br />

j=1<br />

The constants a ij , τ i and b i define q + 1 quadrature rules, where τ i are the<br />

quadrature points and b i and a ij are the quadrature weights for the approximations<br />

<strong>of</strong> the integrals <strong>of</strong> ψ over the interval [0, 1] and [a, b], respectively. Each set <strong>of</strong> those<br />

constants define a RK method and it is customary to depict them in a tabular form<br />

known as Butcher’s table (Table 5.1). It is noted that the index i refers generally


CHAPTER 5. TIME DISCRETIZATION 119<br />

to the stages <strong>of</strong> the RK method.<br />

a 11 a 12 . . . a 1q τ 1<br />

a 21 a 22 . . . a 2q τ 2<br />

. . . .<br />

a q1 a q2 . . . a qq τ q<br />

b 1 b 2 . . . b q<br />

Table 5.1: Butcher’s table<br />

Discretizing the interval [a, b] into N equally spaced elements <strong>of</strong> size h = b−a<br />

N ,<br />

the nodal coordinates are equal to t n = a + nh, n = 0, . . . , N. Let y n be the solution<br />

given by a RK method as t n . Furthermore, as it has been mentioned that a RK<br />

method during each stage computes progressively a better approximation <strong>of</strong> the<br />

solution’s slope, the following constant is defined:<br />

t n,i = t n + τ 1 h, i = 1, . . . , q, (5.2)<br />

which simply expresses the parameterization <strong>of</strong> the sub interval [t n , t n+1 ] to the RK<br />

stages.<br />

Integrating Eq. (5.1) over [t n , t n+1 ], follows:<br />

y ( t n,i) − y (t n ) =<br />

∫ t n +τ i h<br />

which after some algebra may be written as:<br />

y ( t n,i) = y (t n ) +<br />

∫ τi<br />

0<br />

t n<br />

f (t, y (t)) dt,<br />

f (t + sh, y (t n + sh)) ds. (5.3)<br />

Applying the quadrature formulae with constants a ij and τ j , the approximation<br />

y n,i <strong>of</strong> y (t n,i ) results to be equal to:<br />

q∑<br />

y n,i = y n + h a ij f ( t n,j , y n,j) , i = 1, . . . , q. (5.4)<br />

j=1


CHAPTER 5. TIME DISCRETIZATION 120<br />

Integrating now Eq. (5.1) from t n to t n+1 , by applying the variable transformation<br />

t = t n + hs, and approximating the final integral over the interval [0, 1]<br />

with the quadrature rule defined by the points τ i and the weights b i , the following<br />

approximation y n+1 <strong>of</strong> y (t n+1 ) follows:<br />

q∑<br />

y n+1 = y n + h b i f ( t n,i , y n,i) . (5.5)<br />

i=1<br />

As a result, the RK method described by Butcher’s table produces a sequence<br />

<strong>of</strong> approximations given by the following relations:<br />

y 0 = y n = y 0 ,<br />

y n,i = y n + h ∑ qa ij f ( t n,j , y n,j) , 1 ≤ i ≤ q,<br />

j=1<br />

q∑<br />

y n+1 = y n + h b i f ( t n,i , y n,i) .<br />

i=1<br />

(5.6)<br />

In the present implementation <strong>of</strong> the DG discretization <strong>of</strong> the system <strong>of</strong> the<br />

compressible flow equations, explicit time advancement <strong>of</strong> the solution is performed<br />

with high-order strong stability preserving (SSP) Runge-Kutta methods, that preserve<br />

the monotonicity <strong>of</strong> the spatial discretization in any norm or semi-norm coupled<br />

with first-order forward Euler time stepping [85]. That is, if the monotonicity<br />

requirement:<br />

||u n+1 || ≤ ||u n ||, (5.7)<br />

is met for a step size <strong>of</strong> ∆t F E with a first-order forward Euler time stepping, then a<br />

high-order Runge-Kutta method is SSP if there exists a number a ∈ (0, 1] for which<br />

the step size ∆t under ∆t ≤ a∆t F E satisfies the monotonicity requirement.<br />

Extensive use <strong>of</strong> the explicit third order accurate SSP Rung-Kutta method


CHAPTER 5. TIME DISCRETIZATION 121<br />

proposed by Shu and Osher [130] is made:<br />

u (1) = u n + ∆tR(u n ),<br />

u (2) = 3 4 un + 1 4 u(1) + 1 4 ∆tR(u(1) ),<br />

(5.8)<br />

u n+1 = 1 3 un + 2 3 u(2) + 2 3 ∆tR(u(2) ),<br />

where R denotes the residual <strong>of</strong> the system after the DG discretization. Furthermore,<br />

SSP explicit Runge-Kutta methods provided by the PETSc library package are<br />

employed [58, 59].<br />

5.2 Implicit time marching<br />

By grouping together the time dependent and spatial contributions resulting from<br />

the DG discretization <strong>of</strong> either the Euler or the Navier-Stokes equations given in<br />

Eqs. (4.6) and (4.35) respectively, the following system <strong>of</strong> ODEs results:<br />

M · dU dt<br />

+ R(U) = 0, (5.9)<br />

where M is the mass matrix given by Eqs. (4.22) and (4.24) and R is the residual<br />

including all the spatially discretized terms.<br />

Considering for simplicity the backward implicit Euler time discretization <strong>of</strong><br />

Eq. (5.9), obtain:<br />

M · U n+1 − U n<br />

δt<br />

+ R(U n+1 ) = 0. (5.10)<br />

This is a system <strong>of</strong> nonlinear equations <strong>of</strong> the form:<br />

F (U) = 0, (5.11)


CHAPTER 5. TIME DISCRETIZATION 122<br />

and must be solved numerically using a Newton like method that is described next.<br />

5.3 Newton’s method<br />

The Newton iteration for Eq. (5.11) is derived from a multidimensional Taylor series<br />

expansion about a point U k :<br />

F (U k+1 ) = F (U k ) + F ′ (U k ) · (U k+1 − U k ) + higher order terms. (5.12)<br />

Neglecting higher order and setting the right hand side equal to zero, yields Newton’s<br />

method, which basically is an iterative procedure over a sequence <strong>of</strong> linear systems:<br />

J (U k ) · ∆U k = −F (U k ) , U k+1 = U k + ∆U k , k = 0, 1, . . . , (5.13)<br />

given an initial guess for the solution U 0 , J ≡ F ′<br />

= ∂F<br />

∂u<br />

being the approximate<br />

Jacobian <strong>of</strong> the system, U is the state vector to be found and k is the nonlinear<br />

iteration index. The Newton iteration is terminated based on the drop in the norm<br />

<strong>of</strong> the nonlinear function F under a specified tolerance or a sufficiently very small<br />

Newton correction ∆U.<br />

The sequence <strong>of</strong> linear systems is almost always solved<br />

using an iterative method belonging to the class <strong>of</strong> Krylov subspace methods.<br />

Forming each element <strong>of</strong> the Jacobian matrix requires the analytic or discrete<br />

derivatives <strong>of</strong> the system with respect to the state vector U. However, the analytic<br />

determination <strong>of</strong> the derivatives in order to form the Jacobian matrix is error-prone<br />

and even impossible, as the derivatives <strong>of</strong> an nondifferentiable function may need to<br />

be determined. Therefore in this work the Jacobian free method [36, 88] employing<br />

discrete evaluation <strong>of</strong> the Jacobian is used.


CHAPTER 5. TIME DISCRETIZATION 123<br />

5.4 Krylov subspace methods<br />

Krylov subspace methods are approaches for solving large linear systems and first<br />

appeared in the 1950s and became very popular afterward since Reid [123] introduced<br />

them as iterative methods for the solution <strong>of</strong> large linear systems. They can<br />

be considered as projection or generalized projection methods for solving a linear<br />

system A · x = b using the Krylov subspace K j defined as:<br />

K j = span ( r 0 , A · r 0 , A 2 · r 0 , A 3 · r 0 , . . . , A j−1 · r 0<br />

)<br />

,<br />

where r 0 = b − A · x 0 is the system’s residual. These methods require only matrixvector<br />

products during the iteration and this is the key element that favors use <strong>of</strong><br />

Newton’s method in a Jacobian free context as it will be shown.<br />

A wide spectrum <strong>of</strong> iterative methods belongs to the class <strong>of</strong> Krylov methods<br />

[13, 84, 125]. Two primary points <strong>of</strong> classification appear regarding the solution <strong>of</strong><br />

symmetric and non-symmetric systems and the procedure followed for the construction<br />

<strong>of</strong> the bases <strong>of</strong> the Krylov subspace. The bases may be constructed using the<br />

long-recurrence Arnoldi orthogonalization procedure, which generates orthonormal<br />

bases <strong>of</strong> the Krylov subspace. Specifically, the Arnoldi iteration [7] uses the stabilized<br />

Gram-Schmidt process to produce a sequence <strong>of</strong> orthonormal vectors q 1 , q 2 , q 3 , . . .,<br />

known as the Arnoldi vectors such that the vectors q 1 , q 2 , q 3 , . . . , q j span the Krylov<br />

subspace K j . The algorithm for producing the Arnoldi vectors is the following:<br />

1: Start with an arbitrary vector q 1 with norm 1.<br />

2: Repeat for k = 2, 3, . . .<br />

3: q k ← Aq k−1<br />

4: for j to k − 1 do<br />

5: h j,k−1 ← q ∗ j q k<br />

6: q k ← −h j,k−1 q j<br />

7: end for<br />

8: h k,k−1 ← ||q k ||


CHAPTER 5. TIME DISCRETIZATION 124<br />

9: q k ← 1<br />

h k,k−1<br />

q k<br />

The widely used Generalized Minimal RESidual method (GMRES) [126] is an<br />

Arnoldi-based method. In GMRES, the Arnoldi basis vectors form the trial subspace<br />

out <strong>of</strong> which the solution is constructed. One matrix-vector product per iteration<br />

is required to create a new trial vector, and the iterations are terminated based on<br />

a by-product estimate <strong>of</strong> the residual that does not require explicit computation <strong>of</strong><br />

the constructed solution, which is a beneficial feature <strong>of</strong> the algorithm. The residual<br />

minimization property <strong>of</strong> the method is in the Euclidean norm, but requires the<br />

storage <strong>of</strong> all previous Arnoldi basis vectors, leading unfortunately to high requirements<br />

in storage. This necessitates the use <strong>of</strong> efficient preconditioning <strong>of</strong> the linear<br />

system.<br />

5.5 Jacobian-free Newton-Krylov Method<br />

The Jacobian-free Newton-Krylov (JFNK) method is a synergistic combination <strong>of</strong><br />

the Newton method for solving a nonlinear system <strong>of</strong> equations and Krylov subspace<br />

methods for the solution <strong>of</strong> the linear system leading to the Newton correction.<br />

The specific advantage <strong>of</strong> the JFNK method is that the Jacobian <strong>of</strong> the system<br />

is not formed and stored, but only the Jacobian-vector product during the Krylov<br />

iteration is used, which is approximated numerically. This means that the gains<br />

both in processing time and memory requirements are extremely large. However,<br />

the application <strong>of</strong> the Krylov iteration necessitates adequate preconditiong <strong>of</strong> the<br />

linear system to be solved for the successful application <strong>of</strong> the JFNK method.<br />

The origins <strong>of</strong> the JFNK method can be found in efforts motivated by the<br />

solution <strong>of</strong> ODEs and PDEs [30, 31, 33, 56]. The primary goal in all cases has been<br />

the ability to perform a Newton iteration without the need to form the Jacobian.<br />

This has helped the application <strong>of</strong> higher-order implicit integration methods for<br />

ODEs as it is performed in the present work.<br />

In the JFNK approach a Krylov method is used to solve the linear system<br />

resulting from the linearization <strong>of</strong> the initial nonlinear system. An initial residual,


CHAPTER 5. TIME DISCRETIZATION 125<br />

r 0 is defined, given an initial guess for, ∆u 0 , for the Newton correction:<br />

r 0 = −F (u) − J · ∆u 0 (5.14)<br />

Using the GMRES method for the solution <strong>of</strong> the linear system for determining<br />

the Newton correction and denoting by j the GMRES iteration index, the iteration<br />

for the solution <strong>of</strong> the linear system minimizes ‖J · ∆U j + F (U) ‖ 2 within a subspace<br />

<strong>of</strong> small dimensions compared to the total number <strong>of</strong> unknowns. The Newton<br />

correction ∆U j is drawn from the Krylov subspace spanned by the following vectors:<br />

{<br />

r0 , J · r 0 , J 2 · r 0 , . . . , J j−1 · r 0<br />

}<br />

,<br />

and can be written as:<br />

j−1<br />

∑<br />

∆U j = ∆U 0 + β i J i · r 0 , (5.15)<br />

where the scalars β i minimize the residual. Initially the Newton correction ∆U 0 is<br />

set to zero, which asymptotically is a reasonable guess as the approach corrections<br />

should converge to zero at late Newton iterations. Moreover, in practice ∆U j is<br />

determined as a linear combination <strong>of</strong> the orthonormal Arnoldi vectors produced by<br />

GMRES.<br />

Close examination <strong>of</strong> Eq. (5.15) reveals that during the GMRES iterations the<br />

action <strong>of</strong> the Jacobian is taking place in the form <strong>of</strong> matrix-vector products, which<br />

may be approximated as follows [31, 33]:<br />

i=0<br />

J · v ≈<br />

F (U + ɛv) − F (U)<br />

, (5.16)<br />

ɛ<br />

where ɛ is a small perturbation. Eq. (5.16) is simply a first-order Taylor series<br />

expansion approximation to the Jacobian times a vector. For instance, considering<br />

a nonlinear system <strong>of</strong> two equations with two variables, the Jacobian <strong>of</strong> the system<br />

is the following:


CHAPTER 5. TIME DISCRETIZATION 126<br />

J =<br />

⎡<br />

⎢<br />

⎣<br />

∂F 1<br />

∂u 1<br />

∂F 1<br />

∂u 2<br />

∂F 2<br />

∂u 1<br />

∂F 2<br />

∂u 2<br />

JFNK does not require the formation <strong>of</strong> this matrix, instead a vector is formed<br />

that approximates the multiplication <strong>of</strong> this matrix by a vector. Specifically, approximating<br />

F (U + ɛv) − F (U) with a first-order Taylor series expansion about U,<br />

gives:<br />

⎤<br />

⎥<br />

⎦ .<br />

which simplifies to:<br />

F (U + ɛv) − F (U)<br />

ɛ<br />

≈<br />

⎡<br />

⎢<br />

⎣<br />

F 1 (u 1 ,u 2 )+ɛv 1<br />

∂F 1<br />

∂u 1<br />

+ɛv 2<br />

∂F 1<br />

∂u 2<br />

−F 1 (u 1 ,u 2 )<br />

ɛ<br />

F 2 (u 1 ,u 2 )+ɛv 1<br />

∂F 2<br />

∂u 1<br />

+ɛv 2<br />

∂F 2<br />

∂u 2<br />

−F 2 (u 1 ,u 2 )<br />

ɛ<br />

⎤<br />

⎥<br />

⎦ ,<br />

⎡<br />

⎢<br />

⎣<br />

v 1<br />

∂F 1<br />

∂u 1<br />

+ v 2<br />

∂F 1<br />

∂u 2<br />

v 1<br />

∂F 2<br />

∂u 1<br />

+ v 2<br />

∂F 2<br />

∂u 2<br />

⎤<br />

⎥<br />

⎦ = J · v.<br />

It is apparent that the error in the approximation is proportional to ɛ and its evaluation<br />

follows the approach proposed in [126] and [74].<br />

5.6 Preconditioning <strong>of</strong> the JFNK method<br />

Preconditioning in the JFNK method [36, 46, 121, 146] has the purpose <strong>of</strong> reducing<br />

the number <strong>of</strong> GMRES iterations by efficiently clustering close to unity the eigenvalues<br />

<strong>of</strong> the iteration matrix during the solution <strong>of</strong> the linear system. The JFNK<br />

approach has as its main goal the avoidance <strong>of</strong> forming the Jacobian matrix <strong>of</strong> the<br />

system and an effective preconditioner for the linear solution part <strong>of</strong> the method<br />

is necessary. In the present work the preconditioner used is formed by numerically<br />

computing the elements <strong>of</strong> the Jacobian matrix corresponding only to perturbations<br />

<strong>of</strong> the elemental DOFs, using coloring techniques provided by PETSc [58, 59] and


CHAPTER 5. TIME DISCRETIZATION 127<br />

applying an ILU(p) (for p = 0, . . . , 4) factorization procedure for computing the preconditiong<br />

matrix. However, it the experience gained in this work it was observed<br />

that this procedure did not work efficiently especially for large time step sizes.<br />

Unlike the matrix-based GMRES algorithm the Jacobian-free approach enjoys<br />

more flexibility as it allows to approximate the preconditioner matrix without affecting<br />

the quadratic convergence <strong>of</strong> the Newton algorithm. Preconditioning deals<br />

only with the solution <strong>of</strong> the linear system at every Newton iteration and in the<br />

present work advantage <strong>of</strong> this feature was taken by freezing the evaluation <strong>of</strong> the<br />

preconditioner for a number <strong>of</strong> Newton iterations (Jacobian lagging). Significant<br />

efficiency improvements have been reached by the lagging procedure because the<br />

matrix evaluation, for a DG code, is a rather computational expensive task [46].


Chapter 6<br />

Unified limiting<br />

The construction <strong>of</strong> the basis functions and the implementation <strong>of</strong> the DG method<br />

in the computational domain for the standard square element, allows straightforward<br />

numerical treatment <strong>of</strong> mixed-type meshes in a unified way. The expansion<br />

coefficients are defined in the computational space and are unique for every element.<br />

A Taylor series expansion [91] reveals that in the absence <strong>of</strong> discontinuities the solution<br />

coefficients are estimates <strong>of</strong> the solution derivatives. Thus, limiting <strong>of</strong> the<br />

solution coefficients amounts to limiting the solution derivatives.<br />

A significant innovation <strong>of</strong> this work is that a unified procedure was introduced<br />

that allows to preserve monotonicity <strong>of</strong> the solution in a TVB in the mean (TVBM)<br />

sense. A slope limiting procedure, which is transparent to the element type was<br />

developed and explained in this chapter. Specifically, the limiting operations are<br />

performed for every element over the standard square element configuration, where<br />

it is imposed that the slope <strong>of</strong> the solution at the element edges does not exceed the<br />

variation <strong>of</strong> the mean solution across them.<br />

Every element is checked for limiting, separately in each direction, by using<br />

essentially the same TVB limiter proposed by Cockburn and Shu [41, 129]:<br />

m(a 1 , a 2 , . . . , a n ) =<br />

{<br />

a 1<br />

m(a 1 , a 2 , . . . , a n )<br />

if |a 1 | ≤ ML 2 x,y,<br />

otherwise,<br />

(6.1)<br />

128


CHAPTER 6. UNIFIED LIMITING 129<br />

The significant modification <strong>of</strong> the present work is, however, that the parameter<br />

M ≥ 0 for each field is not set to a fixed value, but it is estimated as is shown<br />

bellow. Furthermore, L x,y in Eq. (6.1), is not the rectangular element length (∆x) 2<br />

in the physical space as in the original limiter, but represents the characteristic local<br />

lengths <strong>of</strong> the element in the physical space. The function m(a 1 , a 2 , · · · , a n ), is the<br />

usual minmod function:<br />

⎧<br />

⎨<br />

m(a 1 , a 2 , . . . , a n ) =<br />

⎩<br />

s min<br />

1≤j≤n |a j| if sgn(a 1 ) = sgn(a 2 ) = . . . = sgn(a n ) = s,<br />

0 otherwise.<br />

(6.2)<br />

According to the analysis <strong>of</strong> Cockburn and Shu [40, 129] for scalar conservation<br />

law problems the parameter, M, is an estimate <strong>of</strong> the second derivative <strong>of</strong> the<br />

solution. It is furthermore required that for systems <strong>of</strong> conservation laws to perform<br />

limiting <strong>of</strong> the solution coefficients (solution moments), in the characteristic space.<br />

Evaluation <strong>of</strong> the parameter M is performed for each characteristic field in<br />

the regular computational domain. Limiting for each characteristic variable is performed<br />

separately along each direction. The directional Jacobian, F(u) · n on the<br />

element edges, the corresponding eigenvalues, and eigenvectors are evaluated at the<br />

Roe’s averaged state from the mean interior u − and the exterior u + states in order<br />

to transform the conservative space state variables to the characteristic space<br />

variables. An element is flagged for limiting, if the limiter in Eq. (6.1) returns an argument<br />

other than the first, essentially when the slope is smaller than an appropriate<br />

measure <strong>of</strong> the second derivative value.<br />

For p-adaptive computations <strong>of</strong> arbitrary order <strong>of</strong> approximation P p , p ≥ 1,<br />

only the P 1 expansion <strong>of</strong> the solution is used for elements flagged for limiting in order<br />

to preserve monotonicity. Then, if the solution at an element has been limited, the<br />

approximation order at that element is set to P = 1.<br />

For elements neighboring<br />

limited elements, a gradually higher order <strong>of</strong> approximation could be used until<br />

progressively the order <strong>of</strong> approximation increases to the preset order <strong>of</strong> accuracy<br />

P p .<br />

It is emphasized that for a multistage time marching method and for calcu-


CHAPTER 6. UNIFIED LIMITING 130<br />

lations with a preset order <strong>of</strong> approximation P p , if an element, e, with order <strong>of</strong><br />

approximation p e > 1 is flagged for limiting during a stage, then the solution coefficients<br />

and the residual <strong>of</strong> the discretization corresponding to the higher order part<br />

<strong>of</strong> the approximation are set to zero. The rationale behind this is to keep the order<br />

<strong>of</strong> the approximation strictly equal to P 1 for all limited elements.<br />

For the effective use <strong>of</strong> the TVB limiter, it is crucial to have appropriate<br />

estimates <strong>of</strong> the solution’s second derivative for each field in the characteristic space,<br />

e.g. the parameters M wj , where w i , j = 1, 2, 3, 4 denotes the characteristic fields.<br />

The estimates for the second derivative are given bellow. Furthermore, because<br />

limiting is performed in the computational domain, an estimate <strong>of</strong> the local mesh<br />

size, L x,y , required. In this work L x,y was estimated as suggested by Eq. (3.23).<br />

The value <strong>of</strong> the parameter M (the estimate <strong>of</strong> second derivative <strong>of</strong> the solution<br />

in Eq. (6.1)) is evaluated at every element k, and for each field in the characteristic<br />

space, using central differences <strong>of</strong> the mean solution. Considering the mixed-type<br />

mesh patch given in Fig. 6.1, the parameter M is computed as follows:<br />

Figure 6.1: Mixed-element mesh patch for estimating parameter M.


CHAPTER 6. UNIFIED LIMITING 131<br />

∑<br />

M =<br />

(u e − u k )<br />

∣<br />

∣ , (6.3)<br />

e<br />

where index e counts the neighboring elements <strong>of</strong> k and, which actually is an approximation<br />

<strong>of</strong> the sum <strong>of</strong> the second derivatives <strong>of</strong> the solution. Specifically, performing<br />

a Taylor series expansion [79] <strong>of</strong> the sum given in Eq. (6.3) relative to elements k,<br />

results in:<br />

M = e x u x + e y u y + e xy u xy + e xx u xx + e yy u yy + . . . , (6.4)<br />

where, e x , e y , e xy , . . . , etc are coefficients induced by the local mesh geometry. For a<br />

single type mesh e x = e y = e xy = 0, except for at mesh interfaces. It is observed,<br />

that:<br />

M ≈ u xx + u yy .<br />

In practical computations however, in order to account for the mesh distortion,<br />

truncation error and floating points arithmetic, the following slightly modified value<br />

for the parameter M is used:<br />

ˆM = M +<br />

b<br />

L 2 x,y<br />

, (6.5)<br />

where b is a constant. All results were obtained using the same value b = 0.1.<br />

Therefore in essence in Eq. (6.1), the first condition is replaced by |a 1 | ≤ ML 2 x,y+0.1,<br />

where M is the estimate <strong>of</strong> Eq. (6.3), and L x,y the element size estimate.<br />

6.0.1 Limiting for quadrilateral elements<br />

For a P 1 approximation over a quadrilateral element, the solution expansion is:<br />

u q h = cq 0b q 0 + c q 1b q 1 + c q 2b q 2 + c q 3b q 3. (6.6)<br />

For the standard square element, the bases functions b q j , j = 1, . . . , 4 in Eq. (6.6)<br />

relative to the local Cartesian system (η 1 , η 2 ) <strong>of</strong> the canonical computational domain


CHAPTER 6. UNIFIED LIMITING 132<br />

are:<br />

b q 0 = 1,<br />

b q 1 = η 1 ,<br />

b q 2 = η 2 ,<br />

b q 3 = η 1 η 2 .<br />

Modification <strong>of</strong> the c q j coefficients in Eq. (6.6) is carried out by the TVB limiter as<br />

shown below. It is observed that the base b q 3 is not linear (see Fig. 4.4). Therefore,<br />

the solution coefficient c q 3 could be set to zero if a quadrilateral element is flagged for<br />

limiting. Otherwise, the coefficient c q 3 could be modified with the same procedure<br />

used for the modification <strong>of</strong> the coefficients c q 1 c q 2 considering additional quadrature<br />

points. In this work the c q 3 was set to zero. The solution coefficient c q 1 corresponds to<br />

the u x solution derivative therefore the coefficient c q 1 is limited along the η 1 direction.<br />

For doing so, the P 1 part <strong>of</strong> the solution at the midpoints <strong>of</strong> the edges BC and AD<br />

(see Fig. 4.2) in Ω st is used. The midpoints <strong>of</strong> the edges are chosen for limiting the<br />

solution coefficients, for the following reasons:<br />

ˆ Any ambiguity and closure problems are completely avoided.<br />

ˆ It is computationally efficient.<br />

According to Eq.<br />

solution is equal to:<br />

(6.6), at the midpoints <strong>of</strong> edges BC and AD in Ω st the<br />

U BC = u q (1, 0) − c q 0 = c q 1,<br />

(6.7a)<br />

U AD = u q (−1, 0) − c q 0 = −c q 1,<br />

(6.7b)<br />

and the value for the coefficient c q 1 is expressed in terms <strong>of</strong> the midpoint values as:<br />

c q 1 = U BC − U AD<br />

. (6.8)<br />

2


CHAPTER 6. UNIFIED LIMITING 133<br />

Limiting <strong>of</strong> the solution along the η 1 direction is performed by limiting the midpoint<br />

values U BC and U AD , using Eq. (6.1), as follows:<br />

Ũ BC = m(U BC , c q 0,(i+1,j) − cq 0,(i,j) , cq 0,(i,j) − cq 0,(i−1,j) ),<br />

(6.9a)<br />

Ũ AD = −m(−U AD , c q 0,(i+1,j) − cq 0,(i,j) , cq 0,(i,j) − cq 0,(i−1,j) ),<br />

(6.9b)<br />

where c q 0,(i,j)<br />

denotes the mean solution for the element (i, j) in the computational<br />

domain Ω st . The new limited value <strong>of</strong> c q 1 is computed by substituting in Eq. (6.8)<br />

the limited values from Eqs. (6.9).<br />

Along the η 2 direction the solution coefficient c q 2 is limited using a similar<br />

procedure as for the c q 1 coefficient. Specifically, at the midpoints <strong>of</strong> the edges AB<br />

and CD the solution expansion is equal to:<br />

U AB = u q (0, −1) − c q 0 = −c q 2,<br />

(6.10a)<br />

U CD = u q (0, 1) − c q 0 = c q 2,<br />

(6.10b)<br />

and the value <strong>of</strong> coefficient c q 2 is expressed in terms <strong>of</strong> the midpoint values as:<br />

c q 2 = U CD − U AB<br />

. (6.11)<br />

2<br />

Limiting <strong>of</strong> the values U AB and U CD using Eq. (6.1) is performed as follows:<br />

Ũ AB = −m(−U AB , c q 0,(i,j) − cq 0,(i,j−1) , cq 0,(i,j+1) − cq 0,(i,j) ),<br />

(6.12a)<br />

Ũ CD = m(U CD , c q 0,(i,j) − cq 0,(i,j−1) , cq 0,(i,j+1) − cq 0,(i,j) ).<br />

(6.12b)


CHAPTER 6. UNIFIED LIMITING 134<br />

The new limited value for c q 2 is computed by substituting in Eq. (6.11) the limited<br />

values from Eqs. (6.12).<br />

6.0.2 Limiting for triangular elements<br />

For a P 1 approximation over a triangular element, the solution expansion:<br />

u t h = c t 0b t 0 + c t 1b t 1 + c t 2b t 2, (6.13)<br />

where the basis functions relative to the local Cartesian system (η 1 , η 2 ) <strong>of</strong> Ω st (see<br />

Fig. 4.3 ) are:<br />

b t 0 = 1,<br />

b t 1 − η 2<br />

1 = η 1 ,<br />

2<br />

b t 2 = 3η 2 + 1<br />

.<br />

2<br />

For triangular elements, the coefficients c t 1 and c t 2 correspond to u x and u y ,<br />

respectively. The midpoints <strong>of</strong> the edges are chosen for limiting the solution, and a<br />

similar procedure to that applied for the case <strong>of</strong> quadrilaterals is followed. Specifically,<br />

along the η 1 direction, coefficient c t 1 is limited, by limiting the P 1 part <strong>of</strong><br />

the solution. Then, from the expansion given in Eq. (6.13), the solution at the<br />

midpoints <strong>of</strong> the edges AD and BC is equal to:<br />

U BC = u t (1, 0) − c t 0 = ct 1<br />

2 + ct 2<br />

2 , (6.14a)<br />

U AD = u t (−1, 0) − c t 0 = − ct 1<br />

2 + ct 2<br />

2 , (6.14b)<br />

and coefficient c t 1 is expressed in terms <strong>of</strong> the midpoint values as:


CHAPTER 6. UNIFIED LIMITING 135<br />

c t 1 = U BC − U AD . (6.15)<br />

Limiting <strong>of</strong> the values U AD and U BC is performed using Eq. (6.1) as follows:<br />

Ũ AD = m(U AD , c t 0,(i,j) − c t 0,(i,j−1), c t 0,(i,j+1) − c t 0,(i,j)),<br />

(6.16a)<br />

Ũ BC = m(U BC , c t 0,(i,j) − c t 0,(i,j−1), c t 0,(i,j+1) − c t 0,(i,j)),<br />

(6.16b)<br />

and the new value for c t 1 is computed by substituting in Eq. (6.15) the limited values<br />

from Eqs. (6.16).<br />

Along the η 2 direction the coefficient c t 2 is limited, but the P 1 part <strong>of</strong> the<br />

solution across the edge AB is limited only with the mean solution variation across<br />

that edge. The final value <strong>of</strong> the c t 2 coefficient is:<br />

c t 2 = m(U AB,(i,j) , c t 0,(i,j) − c t 0,(i,j−1), c t 0,(i,j) − c t 0,(i,j−1)). (6.17)<br />

Clearly, the approach described for quadrilateral and triangular elements could<br />

be extended for hexahedral and tetrahedral elements. Again, in three dimensions,<br />

use <strong>of</strong> collapsed coordinates must be employed and limiting will be performed in a<br />

dimension per dimension fashion for the standard cubical element.<br />

Remark 1. As has been noted the parameter M is computed in the characteristic<br />

space. For doing so, first the difference <strong>of</strong> the mean solution is formed at<br />

every element edge and then transformed to the characteristic space.<br />

Remark 2. If the solution at an element is limited then the order <strong>of</strong> accuracy<br />

does not drop to P 0 , since a modified slope is preserved for the approximation.


CHAPTER 6. UNIFIED LIMITING 136<br />

6.1 Positivity preserving limiters for the DG method<br />

The assumption <strong>of</strong> a thermally perfect gas for the equations <strong>of</strong> gas dynamics lead to<br />

the relation between the total energy, pressure and kinetic energy in a flow field given<br />

by Eq. (2.21). It is easily verified that the pressure p is a concave function <strong>of</strong> the state<br />

variables, e.g. defining two states W 1 = [ρ 1 , u 1 , v 1 , E 1 ] T and W 2 = [ρ 2 , u 2 , v 2 , E 2 ] T<br />

it is implied from Jensen’s inequality for 0 ≤ s ≤ 1, that:<br />

p (sW 1 + (1 − s) W 2 ) sp (W 1 ) + (1 − s) p (W 2 ) if ρ 1 , ρ 2 0, (6.18)<br />

and the set <strong>of</strong> admissible states is defined by:<br />

G = {W | ρ, p > 0}, (6.19)<br />

which is convex and this is the key observation for the positivity preserving limiters<br />

for the density and pressure fields developed by Zhang and Shu in [156] where the<br />

solutions coefficients are limited in such a way, so that:<br />

ˆ For smooth solutions accuracy is preserved.<br />

ˆ The scheme remains conservative.<br />

ˆ Limiting for positivity <strong>of</strong> density and pressure is performed locally at every<br />

element.<br />

As in [156] the TVB limiter is applied before the positivity limiters, which will<br />

be described hereafter, but taking into consideration that for the overall success <strong>of</strong> a<br />

positivity preserving scheme, a numerical flux that is preserving the positivity <strong>of</strong> the<br />

variables must be employed, such as the Lax-Friedrichs or the local-Lax-Friedrichs<br />

(Eq. (4.9)). Furthermore, in [112] the positivity <strong>of</strong> an Euler explicit scheme was<br />

analyzed and a sufficient condition was derived for the mean solution <strong>of</strong> the state<br />

variables to be in the set G:


CHAPTER 6. UNIFIED LIMITING 137<br />

∆t<br />

∆x || (|u| + c) || ∞ ≤ aα, (6.20)<br />

where a ∈ (0, 1]. SSP high order Runge-Kutta methods then are suitable for the<br />

positiveness <strong>of</strong> high order time accurate discretizations.<br />

The first step is to limit the coefficients for the density field. This is accomplished by<br />

first computing the minimum value <strong>of</strong> the density ρ min looping over the quadrature<br />

points on the edges. Then the following parameter is defined:<br />

( c<br />

ρ )<br />

0 − ε<br />

θ 1 = min<br />

c ρ , 1 , (6.21)<br />

0 − ρ min<br />

where ε = min (1.0 −13 , c ρ 0, p, ) and p is the mean value <strong>of</strong> the computed pressure.<br />

Then, the final coefficients for the density expansion are modified as follows:<br />

˜c ρ i = θ 1c ρ i , (6.22)<br />

where the index i counts all the bases except for the one corresponding to the mean<br />

solution.<br />

The second step is to limit the pressure. This requires the alteration <strong>of</strong> all the<br />

solution coefficients for every conservative variable. This is accomplished as follows.<br />

Define:<br />

s = (1 − t)W + tq, (6.23)<br />

where W is the mean solution and q is the conservative variables state vector with<br />

the limited density solution. First, calculate:<br />

t = the solution <strong>of</strong> p(s) = ε, if p(q) < ε, (6.24)<br />

then, modify the solution coefficients for each field except for the ones corresponding


CHAPTER 6. UNIFIED LIMITING 138<br />

to the mean solution as:<br />

˜c i = θ 2 c i , (6.25)<br />

where θ 2 is the limiting function given by the following equation:<br />

θ 2 = min (1, t). (6.26)<br />

This limiter was applied only for quadrilateral elements by Zhang and Shu<br />

[156]. In this work, applications for unstructured triangular meshes are demonstrated<br />

with excellent results.


Chapter 7<br />

Numerical results<br />

In the present work a CFD solver based on the previously described DG discretization<br />

was developed and named HoAc as an abbreviation for High Order ACcuracy.<br />

The unified limiting procedure has been implemented in HoAc . Many cases <strong>of</strong> high<br />

speed flows with strong shocks have been analyzed and are presented in this chapter<br />

to demonstrate the improvements obtained with the proposed limiting approach.<br />

The results confirm the outstanding shock capturing and positivity preserving capabilities<br />

<strong>of</strong> the novel limiting approach. Moreover, applications for supersonic inviscid<br />

and subsonic viscous cases will be presented.<br />

In the computations with the new limiting approach, the value M in Eq. (6.1)<br />

was computed at each time step for each characteristic field. The variation <strong>of</strong> the<br />

computed maximum value <strong>of</strong> the Laplacian for each field in the characteristic space:<br />

W 1 ≡ δw 1 = δρ−δp/c 2 , W 2 ≡ δw 2 = −k x δu−k y δv, W 3 ≡ δw 3 = k x δu+k y δv+δp/ρc,<br />

and W 4 ≡ δw 4 = −k x δu − k y δv + δp/ρc, where k = [k x , k y ] is an arbitrary direction<br />

was recorded and will be presented in many cases.<br />

First, verification tests <strong>of</strong> the spatial DG discretization implemented in the<br />

standard space Ω st in HoAc will be presented followed by verification and validation<br />

tests for the limiting procedure, which are performed by comparing the analytical<br />

solution for the standard Sod’s shock tube problem and again for the Sod’s problem<br />

with an extreme pressure ratio. The supersonic flow over a cylinder is considered<br />

139


CHAPTER 7. NUMERICAL RESULTS 140<br />

next employing a P 2 and a P 3 expansion <strong>of</strong> the solution, followed by standard test<br />

cases involving flows with strong shocks proposed in [152]. These standard test cases<br />

for shock capturing schemes include the double mach reflection <strong>of</strong> a Mach 10 strong<br />

shock and a supersonic flow at Mach number <strong>of</strong> 3 in a wind tunnel with a forward<br />

facing step. The Shardin’s problem [60] is solved next, which also confirms the very<br />

good resolution <strong>of</strong> the developed limiting procedure, followed by a test case <strong>of</strong> an<br />

impinging shock at Mach Mach number <strong>of</strong> 5 around the same geometry. Finally, the<br />

diffraction <strong>of</strong> a strong shock over a backward facing step at Mach, Mach number <strong>of</strong><br />

5 is considered for quadrilateral, rectangular and triangular meshes using P 2 and P 3<br />

expansions <strong>of</strong> the solution. Furthermore, it is noted that in all test cases the LLF<br />

flux has been applied and the GLL quadrature points are used for the computation<br />

<strong>of</strong> the integrals. Also, for the application <strong>of</strong> the IC Galerkin projection is employed.<br />

7.1 Solution output<br />

Most present CFD solvers store the solution at the element center or at the nodes in<br />

the mesh. However, a DG method computes the solution in a more general fashion,<br />

as for modal expansions especially the solution coefficients do not have a specific<br />

physical point <strong>of</strong> relevance.<br />

As the solution is discontinuous across the element edges two approaches have<br />

been followed in the present work for outputting and post processing <strong>of</strong> the computed<br />

results: (i) averaging <strong>of</strong> the solution at the nodes <strong>of</strong> the computational mesh<br />

and (ii) a discontinuous output <strong>of</strong> the solution, which is based on outputting the solution<br />

at points over the computational element, whose coordinates are determined<br />

by the inverse transformation <strong>of</strong> equally spaced points in the standard element configuration.<br />

Both approaches give results that are very similar on fine meshes, and for<br />

that, in test cases where fine meshes have been used averaged print <strong>of</strong> the solution<br />

is used.<br />

The discontinuous output <strong>of</strong> the solution leads to a visualization mesh that<br />

is finer than the computational mesh. In Fig. 7.1 a portion <strong>of</strong> a mixed-type ele-


CHAPTER 7. NUMERICAL RESULTS 141<br />

ment computational mesh is shown together with the visualization mesh for a P 4<br />

expansion <strong>of</strong> the solution over every computational element.<br />

Figure 7.1: Computational mesh (blue lines) and visualization mesh (black lines)<br />

for a P 4 expansion <strong>of</strong> the solution over triangular and quadrilateral elements.<br />

In Appendix B the pseudo code for constructing the visualization mesh over a<br />

quadrilateral and a triangular element is provided.<br />

7.2 Code Verification<br />

In this section the verification <strong>of</strong> the implemented DG discretization for the Euler<br />

and the Navier-Stokes equations will be presented, which confirm the correctness <strong>of</strong><br />

the solution algorithms employed in the HoAc solver.<br />

7.2.1 Convergence study for the Euler equations<br />

For verifying the correctness <strong>of</strong> the spatial DG discretization <strong>of</strong> the Euler equations,<br />

a stationary isentropic vortex is considered described by the following IC in the flow<br />

variables:


CHAPTER 7. NUMERICAL RESULTS 142<br />

ρ =<br />

[1 −<br />

] 1<br />

γ−1<br />

(γ − 1) β2<br />

e (1−r2 ) , (7.1)<br />

8γπ<br />

(δu, δv) = β “<br />

2π e 1−r<br />

2 ”<br />

2<br />

[− (y − y 0 ) , (x − x 0 )] , (7.2)<br />

p = ργ<br />

γ . (7.3)<br />

As the above IC satisfy exactly the Euler equations the residual <strong>of</strong> the system in<br />

every field must drop accordingly to the order <strong>of</strong> discretization using a refined mesh.<br />

In Fig. 7.2 the convergence rates for the residual <strong>of</strong> the system corresponding<br />

to the density field is given starting from a P 1 and going up to a P 4 expansion <strong>of</strong> the<br />

solution using the LLF flux. It is observed that the design order <strong>of</strong> the discretization<br />

is achieved.<br />

1<br />

0.01<br />

2<br />

L 2<br />

Error<br />

0.0001<br />

3<br />

P1<br />

P2<br />

P3<br />

P4<br />

1e-06<br />

4<br />

5<br />

1e-08<br />

0.1 1<br />

∆x,∆y<br />

Figure 7.2: Convergence rate for the DG discretization <strong>of</strong> the Euler equations.


CHAPTER 7. NUMERICAL RESULTS 143<br />

7.2.2 Convergence study for the Navier-Stokes equations<br />

The correctness <strong>of</strong> the results for viscous computations is verified by considering the<br />

test case corresponding to the Poiseuille flow, which simply refers to a flow between<br />

two parallel plates under a pressure gradient. An analytical solution <strong>of</strong> this problem<br />

exists under the assumptions <strong>of</strong> incompressible flow and constant viscosity and in<br />

conservative variables is the following:<br />

⎡<br />

U exact =<br />

⎢<br />

⎣<br />

ρ 0<br />

γ−1<br />

⎤<br />

ρ<br />

y (y − b)<br />

0<br />

. (7.4)<br />

(<br />

p0 + dp x) ⎥<br />

+ ρ dx 2 u2 ⎦<br />

1 dp<br />

2µ 0 dx<br />

As HoAc solves the compressible Navier-Stokes equations, the introduction <strong>of</strong> a<br />

source term S to the right hand side <strong>of</strong> the system given in Eq. (2.29) is necessary<br />

in order to retrieve numerical results from the compressible Navier-Stokes equations<br />

corresponding to the incompressible solution given in Eq. (7.4). The source term S<br />

is equal to:<br />

⎡<br />

S =<br />

⎢<br />

⎣<br />

1<br />

2µ<br />

( dp<br />

dx<br />

) 2<br />

[<br />

1<br />

k−1<br />

0<br />

0<br />

0<br />

y (y − b) −<br />

(2y−b)2<br />

2<br />

⎤<br />

. (7.5)<br />

⎥<br />

] ⎦<br />

.<br />

In Fig. 7.3 the convergence rates for the DG discretization <strong>of</strong> the compressible<br />

Navier-Stokes equations using the LDG method are given for a P 1 up to P 4<br />

expansion <strong>of</strong> the solution, with respect to the element length. It is observed that<br />

the design order <strong>of</strong> the method is achieved.


CHAPTER 7. NUMERICAL RESULTS 144<br />

0.1<br />

0.01<br />

L 2<br />

Error<br />

0.001<br />

0.0001<br />

1e-05<br />

1e-06<br />

1e-07<br />

2<br />

3<br />

4<br />

5<br />

P1<br />

P2<br />

P3<br />

P4<br />

1e-08<br />

0.1 1<br />

∆x,∆y<br />

Figure 7.3: Convergence rate for the DG discretization <strong>of</strong> the Navier-Stokes equations<br />

with the LDG method.<br />

7.2.3 Inviscid flow over a cylinder<br />

The potential <strong>of</strong> the discontinuous output <strong>of</strong> the solution is depicted for the computation<br />

<strong>of</strong> the flow over a cylinder at a low Mach number (M ∞ = 0.14). Solutions are<br />

computed with a fixed mixed-type linear element mesh with quadrilateral elements<br />

in the vicinity <strong>of</strong> the cylinder and starting from a P 1 and going up to a P 3 expansion<br />

<strong>of</strong> the solution. It is noted that the curvature based boundary conditions proposed<br />

in [92] have been applied, due to the use <strong>of</strong> straight sided elements, otherwise an<br />

unsteady wake is formed and a steady state solution is not reached.


CHAPTER 7. NUMERICAL RESULTS 145<br />

(a) Computational mesh<br />

(b) P 1 expansion<br />

(c) P 2 expansion<br />

(d) P 3 expansion<br />

Figure 7.4: Mach contours for the inviscid flow over a cylinder using discontinuous<br />

output <strong>of</strong> the solution and the curvature based boundary conditions in [92].<br />

It is observed in Fig. 7.4 that as the order <strong>of</strong> the solution expansion increases so does<br />

the quality <strong>of</strong> the numerical results. Furthermore, in Fig. 7.5 the computed pressure<br />

coefficient at every quadrature point along the edges defining the cylinder geometry<br />

is compared with the analytical solution for the pressure distribution around the<br />

cylinder: C p = 1 − 3 sin 2 (θ). It is clear that as the order <strong>of</strong> the approximation<br />

increases, the numerically evaluated pressure distribution approaches that given by<br />

the analytical solution.


CHAPTER 7. NUMERICAL RESULTS 146<br />

Figure 7.5: Comparison between the numerically evaluated pressure coefficient and<br />

the analytical solution for the inviscid flow around a cylinder.<br />

This test case was solved using an implicit first order backward Euler scheme,<br />

with a step size <strong>of</strong> dt = 1.0 using an unpreconditioned matrix-free approach. The<br />

large number <strong>of</strong> iterations observed in the convergence history plot in Fig. 7.6 for<br />

the density field is attributed to the placement <strong>of</strong> the far field boundary at a distance<br />

<strong>of</strong> 20 unit lengths from the cylinder center.


CHAPTER 7. NUMERICAL RESULTS 147<br />

Figure 7.6: Convergence history for the residual in the density field for the inviscid<br />

flow over a cylinder.<br />

7.2.4 Standard Sod’s shock tube problem<br />

The standard Sod’s shock tube problem with IC [ρ, u, v, p] L = [1, 0, 0, 1] and [ρ, u, v, p] R =<br />

[0.125, 0, 0, 1], which is an exact solution <strong>of</strong> the 1D Euler equations is solved in two<br />

dimensional triangular and quadrilateral meshes, which are shown in Fig. 7.7. This<br />

problem is the first validation test for the developed limiting procedure.


CHAPTER 7. NUMERICAL RESULTS 148<br />

(a) Quadrilateral mesh<br />

(b) Triangular mesh<br />

Figure 7.7: Meshes for the two dimensional Sod’s shock tube problem.<br />

The evolution <strong>of</strong> the limiting procedure for a P 1 numerical solution is depicted<br />

in Fig. 7.8 for the computations performed on the quadrilateral mesh.<br />

0.25<br />

0.2<br />

Time<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

-1 -0.5 0 0.5 1<br />

Length<br />

Figure 7.8: Limited elements in time for the Sod’s shock tube problem using the<br />

quadrilateral mesh <strong>of</strong> Fig. 7.7a.<br />

It is observed that the limiting procedure follows the discontinuities appearing in<br />

this Riemann problem. The contact discontinuity is smeared as time advances and<br />

limiting is performed at the start and end <strong>of</strong> the contact. Similarly, at the start<br />

and end <strong>of</strong> the rarefaction wave the elements are limited. For the smooth parts <strong>of</strong><br />

the flow limiting has been applied in few elements. The final density and entropy<br />

pr<strong>of</strong>iles at time t = 0.25 are shown in Fig. 7.9 and compared with the exact result


CHAPTER 7. NUMERICAL RESULTS 149<br />

for the computations performed on the triangular channel mesh. The numerical<br />

solution was recorded along the center line <strong>of</strong> the channel mesh shown in Fig. 7.7b.<br />

The P 1 computed density distribution <strong>of</strong> Fig. 7.9a agrees quite well with the exact<br />

result. The comparison <strong>of</strong> the most sensitive entropy distribution given in Fig. 7.9b<br />

shows that only small oscillations exist at the location <strong>of</strong> the contact and that the<br />

shock is essentially captured within two elements.<br />

2<br />

1<br />

Exact Solution<br />

Numerical Solution P1<br />

1.8<br />

Numerical Solution P1<br />

Exact Solution<br />

1.6<br />

Density<br />

0.5<br />

Entropy<br />

1.4<br />

1.2<br />

1<br />

0<br />

-1 -0.5 0 0.5 1<br />

Length<br />

(a) Density<br />

-1 -0.5 0 0.5 1<br />

Length<br />

(b) Entropy<br />

Figure 7.9: Comparison <strong>of</strong> the exact density and entropy variation at t = 0.25 with<br />

the two-dimensional numerical solution for the Sod’s shock tube problem in the<br />

triangular channel mesh <strong>of</strong> Fig. 7.7b.<br />

7.2.5 Extreme Sod’s Riemann problem<br />

In order to demonstrate the proper implementation <strong>of</strong> the positivity preserving<br />

limiters a one-dimensional Riemann problem with an exact solution for an extreme<br />

pressure ratio was considered. This problem was solved with a two dimensional<br />

rectangular mesh and it is noted the TVB limiter is applied before the application<br />

<strong>of</strong> the positivity limiters. For this problem, as the solution was advanced in time<br />

it was verified that the positivity preserving limiters were activated. The Sod’s<br />

shock tube problem is considered with the following IC [ρ, r, v, p] l = [1.0, 0, 0, 100],<br />

[ρ, u, v, p] r = [0.125, 0, 0, 0.001], and a specific heat ratio γ = 1.4. For this shock<br />

tube problem an extreme pressure ratio p l /p r = 10 5 exists. In Figs. 7.10, 7.11 and<br />

7.12 the exact and the numerical solution computed with a P 1 expansion and 800


CHAPTER 7. NUMERICAL RESULTS 150<br />

equally sized elements are compared. The computed density, pressure, and entropy<br />

are in good agreement with the exact result.<br />

1<br />

Numeical Solution<br />

Exact Solution<br />

Density<br />

0.5<br />

0<br />

-1 -0.5 0 0.5 1<br />

Length<br />

Figure 7.10: Density plot for the Sod’s problem with a pressure ratio <strong>of</strong> 100000.<br />

100<br />

Numerical Solution<br />

Exact Solution<br />

Pressure<br />

50<br />

0<br />

-1 -0.5 0 0.5 1<br />

Length<br />

Figure 7.11: Pressure plot for the Sod’s problem with a pressure ratio <strong>of</strong> 100000.<br />

100<br />

Entropy<br />

50<br />

Numerical Solution<br />

Exact Solution<br />

0<br />

-1 -0.5 0 0.5 1<br />

Length<br />

Figure 7.12: Entropy plot for the Sod’s problem with a pressure ratio <strong>of</strong> 100000.


CHAPTER 7. NUMERICAL RESULTS 151<br />

For this problem an explicit third-order accurate SSP Runge-Kutta method<br />

was employed with 25 stages provided by the PETSc library package. The time<br />

step used in the computations was equal to dt = 2.0 · 10 −7 and the final time <strong>of</strong> the<br />

simulation was equal to t = 5.0 · 10 −2 . This very low step size was needed as the<br />

maximum Mach number developed in the flow approached the value <strong>of</strong> 12.<br />

7.3 Supersonic flow over a cylinder<br />

The inviscid, supersonic flow at Mach number <strong>of</strong> 2 over a cylinder is considered<br />

with a P 2 and P 3 expansion <strong>of</strong> the solution in order to demonstrate the developed<br />

p-adaptive limiting procedure for arbitrary shaped elements, which is noted again<br />

that it is being performed in the characteristic variable space and over the standard<br />

square element <strong>of</strong> the computational space. Three types <strong>of</strong> meshes quadrilateral,<br />

triangular and a mixed-type element mesh were used for the computations which<br />

are shown in Figs. 7.13, 7.14 and 7.15. An impulsive start <strong>of</strong> the numerical solution<br />

is performed with the following IC: [ρ, u, v, p] = [1.0, 2.0, 0.0, 0.714285714]<br />

Figure 7.13: Quadrilateral mesh for supersonic flow at Mach 2 around a cylinder.


CHAPTER 7. NUMERICAL RESULTS 152<br />

Figure 7.14: Triangular mesh for supersonic flow at Mach 2 around a cylinder.<br />

Figure 7.15: Mixed type mesh for supersonic flow at Mach 2 around a cylinder.<br />

The computed pressure field, which is the same for the numerical solutions obtained<br />

with different meshes, is shown in Fig. 7.16 for a P 3 solution expansion over the<br />

quadrilateral mesh.


CHAPTER 7. NUMERICAL RESULTS 153<br />

Figure 7.16: Pressure contour lines using 30 equally spaced intervals for flow at Mach<br />

2 over a cylinder for the numerical solution obtained with a P 3 approximation.<br />

It is emphasized that once an element is flagged for limiting based on the value<br />

<strong>of</strong> the estimate <strong>of</strong> the Laplacian (second derivative) M, this element is marked as<br />

limited. For a P 2 global expansion the elements next to the limited elements are<br />

P 1 and for the rest <strong>of</strong> the domain are P 2 . For computations with preset higher<br />

order P n expansions and for elements neighboring the P 2 elements the expansion<br />

will be raised to P 3 and progressively to P n . This is demonstrated in Fig. 7.17 for<br />

a computation with P 3 global approximation for a quadrilateral mesh.


CHAPTER 7. NUMERICAL RESULTS 154<br />

Figure 7.17: Limited elements for the flow at Mach 2 over a cylinder for the numerical<br />

solution obtained with a quadrilateral mesh and a P 3 approximation; the blue<br />

elements are limited the adjacent green elements are P 1 expansions, the adjacent<br />

to green red elements are P 2 , and for the rest <strong>of</strong> the domain P 3 expansions are<br />

employed.<br />

The limited elements for a P 2 computation with the triangular and quadrilateral<br />

are shown in Fig. 7.18.


CHAPTER 7. NUMERICAL RESULTS 155<br />

(a) Mixed-type mesh (b) Quadrilateral mesh (c) Triangular mesh<br />

Figure 7.18: Limited elements for the flow at Mach 2 over a cylinder for the numerical<br />

solution obtained with a P 2 approximation; the blue elements are limited<br />

the adjacent red elements are P 1 expansions and for the rest <strong>of</strong> the domain P 2<br />

expansions are employed.<br />

It can be seen that limiting has been applied only for elements neighboring the<br />

shock. For a P 2 global expansion the elements next to the limited elements and for<br />

the rest <strong>of</strong> the domain are P 2 and only the limited elements are P 1 . It is noted<br />

that on the mixed-type element mesh the limited elements are not visible as they<br />

are concentrated in the region where the mesh is very much refined.<br />

A comparison <strong>of</strong> the computed pressure distribution along the stagnation line<br />

obtained with different meshes and polynomial expansions is shown in Fig. 7.19.<br />

The agreement <strong>of</strong> the computations is very good. The computation on the refined<br />

mixed-type element mesh has the sharpest capturing <strong>of</strong> the shock.


CHAPTER 7. NUMERICAL RESULTS 156<br />

Figure 7.19: Comparison <strong>of</strong> the pressure distribution along the stagnation line obtained<br />

from P 2 and P 3 approximations <strong>of</strong> the numerical solutions with different<br />

meshes.<br />

The variation <strong>of</strong> the Laplacian estimate, in essence the parameter M in the<br />

TVB limiter in the current test case, for each field in the characteristic space W i , is<br />

shown in Fig. 7.20 for the P 2 computation using quadrilateral elements.


CHAPTER 7. NUMERICAL RESULTS 157<br />

Figure 7.20: Variation <strong>of</strong> the maximum value <strong>of</strong> parameter M for each characteristic<br />

field obtained during convergence to steady state <strong>of</strong> the numerical solution for the<br />

flow at Mach 2 over a cylinder with a quadrilateral mesh and P 2 solution expansion.<br />

It must be emphasized that until now numerical solutions with the DG method<br />

and the TVB limiters were presented only for rectangular elements and for triangular<br />

elements following the elaborate procedure <strong>of</strong> Cockburn [82], but from the current<br />

test case it is observed that the developed limiting approach works quite well for<br />

elements <strong>of</strong> arbitrary shape, mixed-type element meshes and p-adaptive solutions.<br />

Furthermore, a crisp resolution <strong>of</strong> the shocks is obtained since M < 3 (in contrast to<br />

the suggested value <strong>of</strong> M = 50 by [82]) and fewer elements are limited as depicted<br />

in Fig. 7.18.<br />

7.4 Double Mach Reflection <strong>of</strong> a strong shock<br />

Next, the complex shock pattern generated by the reflection <strong>of</strong> a moving shock at<br />

Mach number <strong>of</strong> 10 inclined at an angle <strong>of</strong> θ = 60 ◦ with respect to a horizontal wall<br />

[152] is solved numerically in order to investigate the potential <strong>of</strong> the p-adaptive


CHAPTER 7. NUMERICAL RESULTS 158<br />

limiting procedure for a flow with a strong moving shock. This is a standard test<br />

problem for shock capturing schemes. It originated by experimental and numerical<br />

studies <strong>of</strong> reflections <strong>of</strong> planar shock waves from wedges. The computational domain<br />

is a rectangle with a width <strong>of</strong> four units and a height <strong>of</strong> one unit. The shock is initially<br />

set up to be inclined at an angle <strong>of</strong> 60 ◦ to the x-axis, and has a Mach number <strong>of</strong><br />

10. Therefore, the IC are: [ρ, u, v, p] = [8.0, 8.25 cos (30 ◦ ) , −8.25 cos (30 ◦ ) , 116.5] for<br />

x < 1 6 + y √<br />

3<br />

and [ρ, u, v, p] = [1.4, 0, 0, 1.0] for x 1 6 + y √<br />

3<br />

. At x = 0 the boundary<br />

conditions are specified to the values corresponding to the post-shock state, and at<br />

x = 4 extrapolation <strong>of</strong> the computed variables is performed. The lower y boundary<br />

for x 1 is a reflecting wall and for x < 1 the fluid values are set by the initial<br />

6 6<br />

post-shock conditions. The upper y boundary is constructed to follow the flow <strong>of</strong> the<br />

diagonal shock such that there is no interaction between the shock and this boundary.<br />

Given the IC here, the intersection <strong>of</strong> the diagonal shock and the upper boundary at<br />

time t occurs at x s (t) = 1 + 1+20t √<br />

6 3<br />

. For x x s (t), the values at the upper y boundary<br />

are set by the initial pre-shock conditions. For x < x s (t), the state variable values<br />

are set by the initial post-shock conditions. In [82, 91, 119] this problem was solved<br />

using a DG discretization <strong>of</strong> the Euler equations on rectangular meshes, but in the<br />

present work is solved using quadrilateral, rectangular and triangular meshes. Both<br />

P 1 and P 2 approximations were used for the numerical solution that was run to a<br />

final time <strong>of</strong> t = 0.2 using the third order explicit SSP Runge-Kutta developed in<br />

[130]. For demonstration, a solution obtained in a relatively coarse quadrilateral<br />

mesh is shown first. The limited elements are shown in Fig. 7.21.<br />

Figure 7.21: Limited elements on the coarse quadrilateral mesh for the double mach<br />

reflection problem at t = 0.2.


CHAPTER 7. NUMERICAL RESULTS 159<br />

It is seen that even for a coarse mesh limiting is applied in a relatively narrow region<br />

around the discontinuities. The final density at t = 0.2 obtained from the numerical<br />

solution with the coarse mesh is shown in Fig. 7.22.<br />

Figure 7.22: Density contours on the coarse quadrilateral mesh for the double mach<br />

reflection problem at t = 0.2.<br />

The numerical solution obtained on a mesh with rectangular elements <strong>of</strong> size<br />

h = 1/240, with a P 1 expansion <strong>of</strong> the solution, for the density field is shown in Fig.<br />

7.23<br />

Figure 7.23: Density contours for the numerical solution computed on the rectangular<br />

mesh, h = 1/240, and a P 1 approximation for the double mach reflection<br />

problem.<br />

and the elements where limiting <strong>of</strong> the solution was performed are shown for time<br />

t = 0.1 and t = 0.2 in Fig. 7.24,


CHAPTER 7. NUMERICAL RESULTS 160<br />

(a) Limited elements at t = 0.1<br />

(b) Limited elements at t = 0.2<br />

Figure 7.24: Limited elements for the numerical solution computed on the rectangular<br />

mesh, h = 1/240, and a P 1 approximation for the double mach reflection. The<br />

blue elements are limited and for the rest <strong>of</strong> the domain a P 1 approximation <strong>of</strong> the<br />

solution is employed.<br />

where the domain decomposition employed for parallel computation is highlighted.<br />

Clearly, seamless limiting is achieved for parallel processing. The computed density<br />

<strong>of</strong> Fig. 7.23 agrees well with other numerical solutions in the literature [82, 91, 119]<br />

obtained with high resolution shock capturing schemes <strong>of</strong> equivalent order.<br />

A numerical solution for an isotropic mesh with triangular elements whose<br />

edge length is equal to h = 1/240 was also obtained for a P 1 solution expansion.<br />

The limited elements at time t = 0.1 and t = 0.2, which are marked in Fig. 7.25,<br />

show that limiting is performed at narrow zones around the discontinuities. The<br />

numerical solution for the density field computed on the triangular mesh is shown<br />

in Fig. 7.26.


CHAPTER 7. NUMERICAL RESULTS 161<br />

(a) Limited elements at t = 0.1<br />

(b) Limited elements at t = 0.2<br />

Figure 7.25: Limited elements for the numerical solution computed on the triangular<br />

mesh, h = 1/240, and a P 1 approximation for the double mach reflection. The blue<br />

elements are limited and for the rest <strong>of</strong> the domain a P 1 approximation <strong>of</strong> the<br />

solution is employed.<br />

Figure 7.26: Density contours for the numerical solution computed on the triangular<br />

mesh, h = 1/240, and a P 1 approximation for the double mach reflection problem.<br />

This test case was also run on the same triangular mesh using a P 2 expansion<br />

<strong>of</strong> the solution. The density field at t = 0.2 is shown in Fig. 7.27 which agrees very<br />

well with the P 1 solution obtained with the rectangular mesh and with the results<br />

given in [119, 152].


CHAPTER 7. NUMERICAL RESULTS 162<br />

Figure 7.27: Density contours for the numerical solution computed on the triangular<br />

mesh, h = 1/240, and a P 2 approximation for the double mach reflection problem.<br />

The variation <strong>of</strong> the parameter M in the application <strong>of</strong> the TVB limiter for<br />

the P 2 solution on the isotropic triangular mesh is depicted in Fig. 7.28, and it is<br />

observed a wide variation <strong>of</strong> its values for each characteristic field.


CHAPTER 7. NUMERICAL RESULTS 163<br />

Figure 7.28: Variation <strong>of</strong> the maximum value <strong>of</strong> parameter M for each characteristic<br />

field obtained during the time advancement <strong>of</strong> the numerical solution for the double<br />

mach reflection problem obtained with a uniform triangular element mesh with a<br />

P 2 approximation and h = 1/240.<br />

7.5 Inviscid flow at Mach number <strong>of</strong> 3 in a tunnel<br />

with a step<br />

The inviscid supersonic flow at Mach number <strong>of</strong> 3 in a tunnel with a step has been<br />

computed with rectangular, triangular and mixed-type element meshes in order to<br />

demonstrate the flexibility and potential <strong>of</strong> the proposed limiting approach. This<br />

is another standard test case for shock capturing schemes and was introduced by<br />

Emery in [51]. The wind tunnel is one unit length wide and three unit lengths


CHAPTER 7. NUMERICAL RESULTS 164<br />

long. The step is 0.2 unit lengths high and is located 0.6 length units from the<br />

left-hand end <strong>of</strong> the tunnel. The corner <strong>of</strong> the step is the center <strong>of</strong> a rarefaction fan<br />

and hence is a singular point <strong>of</strong> the flow field which is seriously affected by large<br />

numerical errors generated just in the neighborhood <strong>of</strong> this singular point. This<br />

may be alleviated in two ways: (i) by using a fine mesh in the region close to the<br />

corner or (ii) by applying special boundary conditions near the corner <strong>of</strong> the step as<br />

demonstrated by Woodward and Collela in [152]. Specifically, the density is reset<br />

so that the entropy has the same value in the zone just to the left and below the<br />

corner <strong>of</strong> the step and the magnitudes <strong>of</strong> the velocities are reset also, but not their<br />

direction, so that the sum <strong>of</strong> enthalpy and kinetic energy per mass has the same<br />

value as in the same zone used to set the entropy. In the present work the first<br />

method is followed.<br />

The numerical solution with the following IC [ρ, u, v, p] = [1.4, 3, 0, 1.0] for the<br />

tunnel with a step was obtained with a P 1 and P 2 expansions, because for computations<br />

with higher order expansions severe time step limitations are encountered<br />

due to the very small elements that had to be used at the corner in order to reduce<br />

the artificial entropy layers generated by the corner singularity.<br />

The first computation was performed with a P 1 expansion <strong>of</strong> the solution for<br />

a mesh with rectangular elements with the edge length equal to h = 1/80. In Figs.<br />

7.29 and 7.30 the limited elements and the numerical solution for the density field<br />

are shown at t = 2.0, respectively.<br />

Figure 7.29: Limited elements for the numerical solution obtained on a rectangular<br />

mesh and P 1 approximation and h = 1/80 for the flow at Mach number <strong>of</strong> 3 in a<br />

wind tunnel with a forward facing step at t = 2.0.


CHAPTER 7. NUMERICAL RESULTS 165<br />

Figure 7.30: Density field for the numerical solution obtained on a rectangular mesh<br />

and P 1 approximation and h = 1/80 for the flow at Mach number <strong>of</strong> 3 in a wind<br />

tunnel with a forward facing step at t = 2.0.<br />

The limited elements <strong>of</strong> Fig. 7.29 closely match the captured discontinuities <strong>of</strong> Fig.<br />

7.30. At later times <strong>of</strong> the simulation the artificial entropy layer generated by the<br />

singularity at the corner creates a Mach stem. In an effort to alleviate the error by<br />

the singularity at the corner, a numerical solution on a isotropic fine triangular mesh<br />

with an average edge length <strong>of</strong> h = 1/100 and for a P 2 expansion <strong>of</strong> the solution<br />

was performed. The limited elements for this numerical solution along with the<br />

computed density field at t = 4 are shown in Figs. 7.31 and 7.32:<br />

Figure 7.31: Limited elements for the numerical solution obtained on a fine triangular<br />

mesh and a P 2 approximation and h = 1/100 for the flow at M = 3.0 in a<br />

wind tunnel with a forward facing step at t = 4.0. Blue elements: limited solution.<br />

Green elements P 1 expansion. Red elements P 2 expansion.


CHAPTER 7. NUMERICAL RESULTS 166<br />

Figure 7.32: Density field for the numerical solution obtained on a fine triangular<br />

mesh and P 2 approximation and h = 1/100 for the flow at Mach number <strong>of</strong> 3 in a<br />

wind tunnel with a forward facing step at t = 4.0.<br />

It is observed that the Mach stem still appears in the flow field. This lead to a third<br />

effort for a numerical solution on a mixed-type element mesh refined at the corner<br />

with triangular elements as shown in Fig. 7.33.<br />

Figure 7.33: Mixed-type element mesh for the inviscid flow at Mach number <strong>of</strong> 3 in<br />

a tunnel with a step.<br />

The average edge length size in the quadrilateral region <strong>of</strong> the mesh is h = 1/40 and<br />

in the triangular region the edge length is equal to h = 3.0 · 10 −3 . The computed<br />

density field at t = 4 is shown in Fig. 7.34. It is observed that the Mach stem has<br />

diminished considerably.


CHAPTER 7. NUMERICAL RESULTS 167<br />

Figure 7.34: Density field for the numerical solution obtained on a mixed-type<br />

element mesh and a P 1 approximation for the flow at Mach number <strong>of</strong> 3 in a wind<br />

tunnel with a forward facing step at t = 4.0.<br />

Furthermore, in Fig. 7.35 the limited elements on the mixed-type element<br />

mesh computation are shown at t = 0.1, where the TVB limiter has smoothly ”fireup”<br />

at the narrow region <strong>of</strong> the developing shock wave in the part <strong>of</strong> the mesh<br />

covered by quadrilateral and triangular elements.<br />

Figure 7.35: Limited elements at t = 0.1 using a mixed-type element mesh for the<br />

inviscid flow at Mach number <strong>of</strong> 3 in a wind tunnel with a forward facing step.<br />

The variation <strong>of</strong> the parameter M versus the number <strong>of</strong> time steps is shown<br />

in Fig. 7.36 where it is observed that there is a wide variation <strong>of</strong> its value for each


CHAPTER 7. NUMERICAL RESULTS 168<br />

characteristic field and differs considerably from the constant value <strong>of</strong> M = 50 as<br />

suggested in [82].<br />

Figure 7.36: Variation <strong>of</strong> the parameter M for the TVB limiter applied on the mixedtype<br />

element mesh for the inviscid flow at Mach number <strong>of</strong> 3 in a wind tunnel with<br />

a forward facing step.<br />

7.6 Diffraction <strong>of</strong> a strong shock over a backward<br />

facing step<br />

The effectiveness <strong>of</strong> the positivity preserving limiters combined with the TVB limiter<br />

for shock capturing is tested for the diffraction <strong>of</strong> a Mach number <strong>of</strong> 5.09 right<br />

moving shock over a backward facing step. This problem is solved on meshes with<br />

rectangular, quadrilateral and triangular elements. The IC for this problem is a<br />

right moving shock wave. The element size for the solutions was h = 0.025 for the<br />

rectangular and triangular meshes and <strong>of</strong> comparable size for the region close to the


CHAPTER 7. NUMERICAL RESULTS 169<br />

corner for the mesh with quadrilateral elements. Sample meshes along with contour<br />

lines for the density field at t = 0.8 <strong>of</strong> the diffraction <strong>of</strong> the shock wave are given in<br />

Fig. 7.37.<br />

(a) Quadrilateral elements mesh<br />

(b) Rectangular elements mesh<br />

(c) Triangular elements mesh<br />

Figure 7.37: Sample meshes with density contours at t = 0.8 for the diffraction<br />

<strong>of</strong> a strong shock over a backward facing step using a P 2 solution expansion. The<br />

black square shows where the computational domain was clipped for showing the<br />

underlying computational mesh.


CHAPTER 7. NUMERICAL RESULTS 170<br />

A P 2 expansion <strong>of</strong> the solution has been used in all cases. For demonstration,<br />

a solution with P 3 expansion was also computed. The elements for which limiting<br />

was applied at time t = 2.0 for the computations performed with different meshes<br />

are shown in Fig. 7.38.<br />

(a) Quadrilateral elements<br />

(b) Rectangular elements<br />

(c) Triangular elements<br />

Figure 7.38: Limited elements for the diffraction <strong>of</strong> strong shock over a backward<br />

facing step at t = 2.0. Blue elements: limited solution. Green elements P 1 expansion.<br />

Red elements P 2 expansion.<br />

Comparing Fig. 7.38 with the computed density at the same time in Fig. 7.39:


CHAPTER 7. NUMERICAL RESULTS 171<br />

(a) Quadrilateral elements<br />

(b) Rectangular elements<br />

(c) Triangular elements<br />

Figure 7.39: Density contours for the diffraction <strong>of</strong> strong shock over a backward<br />

facing step at t = 2.0 with 30 equally spaced intervals from 0.06 to 7.16.<br />

it is seen that limited elements were obtained at the shock locations. Comparing<br />

Fig. 7.39 with the pressure distribution in Fig. 7.40:


CHAPTER 7. NUMERICAL RESULTS 172<br />

(a) Quadrilateral elements<br />

(b) Rectangular elements<br />

(c) Triangular elements<br />

Figure 7.40: Pressure contours for the diffraction <strong>of</strong> strong shock over a backward<br />

facing step at t = 2.0 with 30 equally spaced intervals from 0.09 to 31.9.<br />

the location <strong>of</strong> the contact discontinuity can be identified. It appears no limiting was<br />

performed in the regions <strong>of</strong> the contact discontinuities. The line plot <strong>of</strong> Fig. 7.41 for<br />

the density and pressure along a line starting from the left side <strong>of</strong> the computational<br />

domain, ending at its right side and being above the step, also demonstrates the<br />

existence and quite good capturing <strong>of</strong> the contact discontinuity, which is further<br />

justified by the numerically computed gradient <strong>of</strong> pressure and density in Figs. 7.42<br />

and 7.43.


CHAPTER 7. NUMERICAL RESULTS 173<br />

30<br />

25<br />

Density<br />

Pressure<br />

Density, Pressure<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 2 4 6 8 10 12<br />

Length<br />

Figure 7.41: Plot <strong>of</strong> density and pressure along the line defined by points [0, 7] and<br />

[13, 7].<br />

Figure 7.42: Numerically computed pressure (left) and density (right) gradient field<br />

for the diffraction <strong>of</strong> strong shock over a backward facing step at t = 2.0, using a<br />

rectangular elements mesh.


CHAPTER 7. NUMERICAL RESULTS 174<br />

Figure 7.43: Numerically computed pressure (left) and density (right) gradient field<br />

for the diffraction <strong>of</strong> strong shock over a backward facing step at t = 2.0, using a<br />

triangular elements mesh.<br />

The variation <strong>of</strong> the maximum value <strong>of</strong> the parameter M is shown in Fig.<br />

7.44 versus the iteration number for the simulation on the mesh with rectangular<br />

elements and a P 2 expansion.<br />

30<br />

W 1<br />

25<br />

W 3<br />

W 3<br />

W 4<br />

20<br />

M 15<br />

10<br />

5<br />

0<br />

0 5000 10000<br />

Iteration<br />

Figure 7.44: Variation <strong>of</strong> the second derivative estimate for the diffraction <strong>of</strong> a strong<br />

shock over a backward facing step using a P 2 expansion on a rectangular elements<br />

mesh.


CHAPTER 7. NUMERICAL RESULTS 175<br />

It is observed that there is a wide variation <strong>of</strong> the maximum value <strong>of</strong> the parameter<br />

M for each field as the solution advances in time. In contrast to [156] where fixed a<br />

high value <strong>of</strong> the parameter M was prescribed (M = 100), with the present limiting<br />

approach these values were evaluated during the time advancement <strong>of</strong> the solution<br />

and limiting was performed only where the variation <strong>of</strong> the computed solution was<br />

very large. These regions are most <strong>of</strong> the time shock waves.<br />

Additional computations on the rectangular mesh were performed with P 1 and<br />

P 3 expansions <strong>of</strong> the solution, in order to examine the effect on the results <strong>of</strong> the<br />

combined application <strong>of</strong> the shock capturing and positivity preserving limiters. In<br />

Fig. 7.45 a plot over a line defined by the points [2, 5.5] and [5, 4] at time t = 0.9 <strong>of</strong><br />

density and pressure is shown.<br />

14<br />

12<br />

Pressure P2<br />

Density P1<br />

Denstiy P2<br />

Density P3<br />

Density, Pressure<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0.5 1 1.5 2 2.5<br />

Length<br />

Figure 7.45: Plot <strong>of</strong> density and pressure along a line defined by the points [2, 5.5]<br />

and [5, 4] at time t = 0.9.<br />

Due to the fine mesh and the non-limiting <strong>of</strong> the solution at the slip line as it is<br />

evident for the P 2 computation from Fig. 7.38, the resolution <strong>of</strong> the computations<br />

is the same. However, the location <strong>of</strong> the shock has been captured in a different<br />

position for the P 1 computation in comparison with the P 2 and P 3 computations,<br />

which actually for the last two is the same.<br />

The large-scale wiggles at time t = 2.0 <strong>of</strong> Fig. 7.39(a) are due to the irregular<br />

and coarse mesh since the quadrilaterals expand away from the corner. The enlarged


CHAPTER 7. NUMERICAL RESULTS 176<br />

plots <strong>of</strong> Fig. 7.37 also show that some wiggles could be introduced from irregular<br />

meshes (see Fig. 7.39(a)). For the finer and regular rectangular and triangular<br />

meshes <strong>of</strong> Figs. 7.39(b),(c) the wiggles are not caused by the mesh or the TVB<br />

limiter. Note that in Fig. 7.37 the slope limiter and the positivity preserving<br />

limiter have been applied already many times until t = 0.9, but no wiggles in<br />

the computed solution appear. The wiggles at the end <strong>of</strong> the simulation at time<br />

t = 2.0 are caused by small scale disturbances that are generated in the region<br />

between the shock and the contact discontinuity, and they are not well resolved.<br />

In the absence <strong>of</strong> physical viscosity, these disturbances cannot diffuse. Depending<br />

on the numerical viscosity introduced by the numerical scheme, and the resolution<br />

<strong>of</strong> the mesh, some disturbances may dissipate and the rest appear as wiggles in<br />

the computed density and pressure plots. The computed vorticity field at time<br />

t = 2.0, shown bellow in Fig. 7.46, for the simulations performed on the rectangular<br />

and triangular meshes, demonstrate that for the rectangular element mesh these<br />

disturbances are concentrated in the region 1.5 < x < 9 marked by a red line<br />

in Fig. 7.46(a). For the computation with the triangular mesh, which provides<br />

higher resolution per unit area, the onset <strong>of</strong> a second contact discontinuity after<br />

the Mach stem is evident. Further away, this contact could not be resolved in the<br />

computations. For the part <strong>of</strong> the domain were the contact remains unresolved (the<br />

region 5.5 < x < 11 also marked by a red line in Fig. 7.46(b)) wiggles are found in<br />

the computed solution.


CHAPTER 7. NUMERICAL RESULTS 177<br />

(a) Rectangular elements<br />

(b) Triangular elements<br />

Figure 7.46: Vorticity field for the diffraction <strong>of</strong> a Mach 5 shock over a backward<br />

facing step.


CHAPTER 7. NUMERICAL RESULTS 178<br />

intervals<br />

7.7 Shardin’s problem<br />

The effectiveness <strong>of</strong> the novel p-adaptive limiting procedure is further examined for<br />

the standard Schardin’s problem [60] in order to validate the use <strong>of</strong> the limiting<br />

procedure for shock capturing. For this problem a right moving shock at a Mach<br />

number <strong>of</strong> 1.34 impinges on a triangle. An unsteady flow is produced accompanied<br />

by the generation <strong>of</strong> vortices and complex wave patterns resulting from the diffraction<br />

<strong>of</strong> the shock at the triangle’s corners and multiple interactions between the<br />

reflected shock waves and the generated vortices. The complete problem setup and<br />

the experimental results are described in detail in [34].<br />

For this problem two types <strong>of</strong> meshes have been used with different orders <strong>of</strong><br />

expansion for the solution. In order to reduce the problem size only half <strong>of</strong> the<br />

triangle was considered and symmetry conditions were imposed. An unstructured<br />

triangular mesh with a P 1 expansion and approximately 200000 elements and a<br />

quadrilateral mesh with a P 2 expansion and approximately 285000 elements. Both<br />

meshes have an average mesh length <strong>of</strong> h = 0.002. The reason for employing a P 1<br />

expansion on the triangular mesh is that triangles provide almost double resolution<br />

over the same area in the computational domain compared to quadrilateral elements.<br />

The computed density and pressure fields at time t = 128µs are given in<br />

Figs. 7.47 and 7.48. Indeed the computed density field by both triangular and<br />

quadrilateral mesh numerical solutions with P 1 and P 2 expansions respectively are<br />

<strong>of</strong> comparable resolution.


CHAPTER 7. NUMERICAL RESULTS 179<br />

(a) P 1 expansion. Triangular elements<br />

(b) P 2 expansion. Quadrilateral elements<br />

Figure 7.47: Density contours for Schardin’s problem with 30 equal spaced intervals<br />

from 0.6 to 2.4 using a P 2 solution expansion


CHAPTER 7. NUMERICAL RESULTS 180<br />

(a) P 1 expansion. Triangular elements<br />

(b) P 2 expansion. Quadrilateral elements<br />

Figure 7.48: Pressure contours for Schardin’s problem with 30 equal spaced intervals<br />

from 0.4 to 2.14 using a P 2 expansion.<br />

It is observed that both expansion sets and mesh types reveal the same structure <strong>of</strong><br />

the flow field behind the triangle, with the exception, that the results produced by<br />

the P 2 expansion capture each discontinuity in a narrower region. Furthermore, in<br />

Fig. 7.49 where the numerical Schlieren for both polynomial expansions is shown:


CHAPTER 7. NUMERICAL RESULTS 181<br />

(a) P 1 expansion. Triangular elements<br />

(b) P 2 expansion. Quadrilateral elements<br />

Figure 7.49: Numerical Schlieren for Schardin’s problem using a P 2 solution expansion.<br />

it can be observed that the complex wave pattern produced by the interaction <strong>of</strong><br />

the vortices and the impinging and reflected shock waves is basically the same,<br />

but that <strong>of</strong> the P 2 expansion has resolved better the vortices as also the diffracted<br />

waves and the slip lines. However, in the P 2 computation wiggles appear at the slip


CHAPTER 7. NUMERICAL RESULTS 182<br />

lines locations, but not for the P 1 computation. This is due to the fact that the<br />

combination <strong>of</strong> a fine mesh and higher order expansion lead to smaller numerical<br />

viscosity. This observation is consistent with the trends found in high resolution<br />

inviscid flow computations carried out with increasing order <strong>of</strong> accuracy and mesh<br />

resolution using DG [82] and WENO discretizations [12]. The numerical results<br />

presented are in very good agreement with the experimental results given in [34],<br />

where the shadowgraph at t = 128µs shown in Fig. 7.50:<br />

Figure 7.50: Shadowgraph for the Schardin’s problem at t = 128µs.<br />

compares very well with the numerical results depicted previously in Fig. 7.49,<br />

and the numerical results obtained by a high fidelity numerical simulation for this<br />

problem performed by Chaudhuri et al. [35] obtained using a WENO scheme on a 8.4<br />

million points mesh. However, the P 2 results in Fig. 7.49 which were obtained with<br />

the 285000 element mesh required approximately 20 CPU hours on 50 processors<br />

in contrast to the results <strong>of</strong> [35] that required approximately 500 CPU hours using<br />

20-40 processors.<br />

In order to demonstrate that the accuracy is preserved in smooth regions <strong>of</strong> the<br />

flow field for the Schardin’s problem with the present shock capturing approach, a


CHAPTER 7. NUMERICAL RESULTS 183<br />

grid and p-refinement study was performed on the triangular element mesh. Specifically,<br />

at the region behind the triangle where the vortex is formed, a refinement <strong>of</strong><br />

the mesh was performed. The highly refined mesh and the computed Schlieren with<br />

P 2 expansion is shown in Fig. 7.51. Clearly, the shear layer is captured with 4 cells,<br />

while the shocks are captured with two cells.<br />

Figure 7.51: Numerical Schlieren for the Shardin’s problem on the finest mesh and<br />

using P 2 solution expansion.<br />

The first mesh refinement performed using an element size equal to h 1 = 0.001<br />

in the refined mesh region, while for the rest <strong>of</strong> the domain the size <strong>of</strong> the elements<br />

was h = 0.002 as before. Numerical solutions were obtained using P 1 and P 2<br />

solution expansions. The numerical Schlieren at time t = 128µs is shown in Fig.<br />

7.52, for both expansions, and it is apparent that a comparison between Figs. 7.52<br />

and 7.49(a) reveals that the solution on the refined mesh has resolved better the<br />

vortex formation behind the triangle. Furthermore, a solution was obtained with a<br />

more refined mesh having an element size <strong>of</strong> h 2 = 0.0005 in the region behind the<br />

triangle. As in the previous mesh refinement study, two numerical solutions were<br />

obtained using P 1 and P 2 solution expansion. The numerical Schlieren at a time<br />

t = 128µs is shown in Fig. 7.53 where it is observed that both solution expansions<br />

have resolved even finer and more complex flow structures.


CHAPTER 7. NUMERICAL RESULTS 184<br />

(a) P 1 expansion.<br />

(b) P 2 expansion.<br />

Figure 7.52: Numerical Schlieren for Schardin’s problem on the refined triangular<br />

mesh with mesh element size in the vortex region equal to h 1 = 0.001.


CHAPTER 7. NUMERICAL RESULTS 185<br />

(a) P 1 expansion.<br />

(b) P 2 expansion.<br />

Figure 7.53: Numerical Schlieren for Schardin’s problem on the refined triangular<br />

mesh with mesh element size in the vortex region equal to h 2 = 0.0005.<br />

7.7.1 Strong Mach 5 shock impingement on a triangle<br />

In this test case the same geometry employed in the computations for the standard<br />

Schardin’s problem is used, but the IC correspond to a right moving shock at a Mach


CHAPTER 7. NUMERICAL RESULTS 186<br />

number <strong>of</strong> 5.09. The triangular mesh has been used but with a P 2 expansion <strong>of</strong> the<br />

solution as higher gradients in the unsteady flow behind the triangle were expected.<br />

Furthermore, the positivity limiters proposed in [156] were employed, otherwise the<br />

computations were not able to advance in time long enough due to the appearance<br />

<strong>of</strong> negative density and pressure.<br />

The computed density field is shown in Fig. 7.54. In Fig. 7.55, the elements<br />

where the solution was limited along with the computed pressure gradient gray scale<br />

contours are shown for three characteristic non dimensional times. The limited<br />

elements are marked with blue color, for elements neighboring limited elements P 1<br />

approximation is used and these elements are marked with green color. For the rest<br />

<strong>of</strong> the domain, marked with red color, P 2 approximation is used. It is observed that<br />

limiting has been performed mostly at the elements where the impinging shocks<br />

(detected by pressure gradients) as well as through elements, which the reflected<br />

shocks are crossing. Moreover, in Fig. 7.55 the numerical Schlieren is shown where<br />

it is evident that limiting has not been performed over the elements where the<br />

contact discontinuities are formed.


CHAPTER 7. NUMERICAL RESULTS 187<br />

(a) t = 0.06<br />

(b) t = 0.08<br />

(c) t = 0.1<br />

Figure 7.54: Density contours using a P 2 expansion for the flow <strong>of</strong> a strong shock<br />

moving at Mach number <strong>of</strong> 5 around a triangle with 30 equal spaced intervals from<br />

0.28 to 13.28. Results are depicted at non-dimensional time.


CHAPTER 7. NUMERICAL RESULTS 188<br />

(a) t = 0.06<br />

(b) t = 0.08<br />

(c) t = 0.1<br />

Figure 7.55: Limited elements for the flow <strong>of</strong> a strong shock moving at Mach number<br />

<strong>of</strong> 5 around a triangle along with the gradient <strong>of</strong> pressure over 30 equally spaced intervals<br />

from 0 to 1000. Results are depicted at non-dimensional time. Blue elements:<br />

limited solution. Green elements P 1 expansion. Red elements P 2 expansion.


CHAPTER 7. NUMERICAL RESULTS 189<br />

7.8 Flat plate boundary layer<br />

The viscous flow over a flat plate with no pressure gradient at M = 0.2 and Re = 10 4<br />

is considered to further validate the viscous calculation with HoAc by comparing the<br />

velocity pr<strong>of</strong>iles for a P 1 , P 2 and a P 3 solution expansion. The free stream conditions<br />

are: [ρ, u, v, p] = [1, 0.2, 0, 0.714285714] with γ = 1 and the computational domain<br />

4<br />

has dimensions <strong>of</strong> [−2, 2] × [0, 10] with half <strong>of</strong> the lower horizontal line representing<br />

a symmetry boundary and the other half an adiabatic wall. At x = −2 the flow<br />

velocities and pressure are specified, while the density is extrapolated from the<br />

interior. At x = 2 and y = 10 all the computed variables are extrapolated except<br />

for the pressure which is specified at its mean stream value. A discontinuous output<br />

<strong>of</strong> the solution is employed as a relatively coarse mesh is used for the computations.<br />

Specifically, quadrilateral elements are used with a height <strong>of</strong> h = 10 −3 for the first<br />

layer <strong>of</strong> elements and a stretching factor <strong>of</strong> 1.2 in the direction normal to the wall.<br />

In Figs. 7.56 and 7.57 the computed velocity pr<strong>of</strong>iles are shown with comparison<br />

to the Blasius solution. It is observed that the P 1 computation hardly agrees with<br />

the Blasius solution especially for the V-velocity pr<strong>of</strong>ile. But, the P 2 computation<br />

is close to analytical solution and the P 3 computation achieves the best resolution<br />

for the boundary layer.


CHAPTER 7. NUMERICAL RESULTS 190<br />

Figure 7.56: Comparison <strong>of</strong> the U-velocity pr<strong>of</strong>ile with the Blasius solution for the<br />

flow over a flat plate.<br />

Figure 7.57: Comparison <strong>of</strong> the V-velocity pr<strong>of</strong>ile with the Blasius solution for the<br />

flow over a flat plate.


CHAPTER 7. NUMERICAL RESULTS 191<br />

The velocity field vectors along with the vorticity field for the P 3 computation<br />

is depicted in Fig. 7.58, where it is observed that the boundary layer thickness<br />

grows along the plate and the velocity pr<strong>of</strong>iles are nicely resolved by the numerical<br />

solution.<br />

Figure 7.58: Velocity field vectors and vorticity field for the flat plate boundary<br />

layer solution using a P 3 solution expansion.<br />

7.9 Unsteady viscous flow over tandem airfoils<br />

This problem was a test case that was solved and presented in the first International<br />

Workshop on High Order CFD Methods [89]. The geometry <strong>of</strong> the problem consists<br />

<strong>of</strong> two tandem NACA 0012 airfoils with the forward airfoil at 10 ◦ angle <strong>of</strong> attack.<br />

The free stream flow conditions correspond to a Mach number M = 0.2 and a<br />

Reynolds number Re = 10 4 . The IC applied to this test case correspond to a C 1<br />

initial solution for the velocity field:<br />

⎧<br />

√ γp∞<br />

⎨<br />

V = (u, v) = M ∞ (1, 0) (<br />

ρ ∞ ⎩ sin<br />

)<br />

πd<br />

2δ 1<br />

1, if d > δ 1 ,<br />

, if d ≤ δ 1 ,<br />

(7.6)<br />

with δ 1 = 0.05 and d being the distance to the closest wall with the density and<br />

pressure initialed to their free stream values. A mixed-type linear element mesh<br />

was generated around the geometry <strong>of</strong> the airfoils with quadrilateral elements in the<br />

region close to the airfoils in order to achieve better resolution for the boundary layer.


CHAPTER 7. NUMERICAL RESULTS 192<br />

In Fig. 7.59 the visualization mesh for a P 4 solution expansion over quadrilateral<br />

and triangular elements is shown.<br />

Figure 7.59: Visualization mesh for a P 4 computation for the two tandem NACA<br />

0012 airfoils.<br />

For this specific problem a sensitivity study regarding the far field boundary<br />

distance was carried out first in order to resolve the lift and drag coefficients within<br />

0.01 counts using a P 2 solution expansion over each element. The far field boundary<br />

was formed by a circle enclosing the airfoils. Three simulations were performed in<br />

order to examine the effect <strong>of</strong> the far field distance from the airfoils. These corresponded<br />

to the following radii <strong>of</strong> the circles: R = 25, R = 50, R = 100 times<br />

the chord length <strong>of</strong> the airfoil, which was assumed equal to unity. The sensitivity<br />

study was performed using an explicit third-order SSP Runge-Kutta scheme with 16<br />

stages up to t = 100, using a time step size <strong>of</strong> dt = 1.0e − 3, corresponding to nondimensional<br />

units. The cost <strong>of</strong> the computations is expressed in work units defined<br />

by comparing the execution time T 2 , excluding the initialization, post-processing<br />

data preparation time and file I/O time, to TauBench time T 1 , which is an unstructured<br />

grid benchmark. The respective kernels are derived from Tau - a Navier-Stokes<br />

solver, which has been developed at DLR in Germany.


CHAPTER 7. NUMERICAL RESULTS 193<br />

Work units = T 2<br />

T 1<br />

.<br />

In Table 7.1 the work units, number <strong>of</strong> elements, number <strong>of</strong> DOF and the L 2 error<br />

for the whole time <strong>of</strong> the simulation for the aerodynamic coefficients for both airfoils<br />

are given. All the simulations were performed using 8 cores. From the results given<br />

in Table 7.1, the distance <strong>of</strong> the far field for the time accurate simulations was chosen<br />

to be R = 50.<br />

R Elem Work Units DOF C l forward C d forward C l aft C d aft<br />

25 2708 1163.49 146496 – – – –<br />

50 2788 1312.3 150336 5.4880e-7 1.0240e-7 2.1443e-6 1.6780e-7<br />

100 2914 1649.78 156384 6.8725e-8 1.0516e-8 1.2237e-7 3.0579e-8<br />

Table 7.1: Results for the sensitivity study regarding the far field boundary distance.<br />

Furthermore, a p-refinement study was performed using a P 2 , P 3 and P 4<br />

solution expansion for a time accurate computation using an explicit third order<br />

SSP Runge–Kutta method with 16 stages, provided by PETSc, and a small step<br />

size equal to dt = 5.0e − 4 for a simulation time equal to t = 100. The time history<br />

for the lift and drag coefficients for the forward and the aft airfoil are given in Figs.<br />

7.60 and 7.61, respectively.


CHAPTER 7. NUMERICAL RESULTS 194<br />

(a) Lift coefficient t = 0 − 100 (b) Lift coefficient t = 40 − 100<br />

(c) Drag coefficient t = 0 − 100 (d) Drag coefficient t = 40 − 100<br />

Figure 7.60: Lift and drag coefficients for the forward airfoil.


CHAPTER 7. NUMERICAL RESULTS 195<br />

(a) Lift coefficient t = 0 − 100 (b) Lift coefficient t = 80 − 100<br />

(c) Drag coefficient t = 0 − 100 (d) Drag coefficient t = 80 − 100<br />

Figure 7.61: Lift and drag coefficients for the aft airfoil.<br />

It is observed that a time periodic response appears to develop for both airfoils after<br />

time t = 80. In Figs. 7.62 and 7.63 the velocity and vorticity contours at t = 80 are<br />

shown for a P 2 and a P 4 solution expansion using a discontinuous output, where<br />

the difference between the resolved flow fields is apparent.


CHAPTER 7. NUMERICAL RESULTS 196<br />

(a) P 2 Solution<br />

(b) P 4 Solution<br />

Figure 7.62: Velocity contours for the flow over NACA0012 tandem airfoils.<br />

(a) P 2 Solution<br />

(b) P 4 Solution<br />

Figure 7.63: Vorticity contours for the flow over NACA0012 tandem airfoils.


Chapter 8<br />

Conclusions and Future Work<br />

The present work developed and evaluated a priori indicators <strong>of</strong> mesh distortion.<br />

This was accomplished by deriving the complete truncation error expression for the<br />

FV discretization <strong>of</strong> the field gradient in two dimensions. The complexity <strong>of</strong> the<br />

mathematical operations involved was overcome via use <strong>of</strong> symbolic mathematics<br />

s<strong>of</strong>tware.<br />

The presented work analyzed a generally-distorted mesh (structured, unstructured<br />

and mixed-type element) into elementary distortions (stretching, skewness,<br />

shearing, rotation), as well as three common types <strong>of</strong> interfaces, and those were directly<br />

related to the TE. The derived analytic expressions are relatively simple and<br />

amenable to future work on improving the mesh. This is especially true for index<br />

Q for which analytic expressions were presented directly relating it to the degree <strong>of</strong><br />

stretching and skewness, as well as to mixed-type element mesh interfaces.<br />

The two indices (Q and q) “followed” the grid distortion quite faithfully with<br />

Q being more focused on these local areas. Index q marks broader areas <strong>of</strong> the mesh<br />

due to the second order error terms included in its definition. It also picks-up small<br />

(background) distortions which may not be <strong>of</strong> interest. Given the fact that Q is<br />

quite simpler to calculate, it is the recommended quality index. A disadvantage <strong>of</strong><br />

Q is that it cannot indicate the quality <strong>of</strong> distorted meshes that still yield second<br />

order accurate computation <strong>of</strong> the gradient. Shearing and rotation do change the<br />

197


CHAPTER 8. CONCLUSIONS AND FUTURE WORK 198<br />

TE (i.e. only index q “picks them” up) but they do not reduce the order <strong>of</strong> accuracy.<br />

However, the goal is to fix meshes that degrade the accuracy to a lower order.<br />

The following observations were also made for both indices: (i) irregularlyshaped<br />

small distortions were captured faithfully, (ii) small distortions were detected<br />

even though much larger ones existed in the mesh, and (iii) distortions were captured<br />

in any direction.<br />

Employment <strong>of</strong> the centroid dual discretization must be avoided on general<br />

structured and mixed-type element meshes due to its inconsistency in the evaluation<br />

<strong>of</strong> the field gradient.<br />

Future work involves two directions: (i) working the analytical expressions for<br />

the three-dimensional case for mixed-type element (structured/unstructured) grids,<br />

and (ii) using the analytic expressions <strong>of</strong> the present work for a priori improvement<br />

<strong>of</strong> the mesh.<br />

In the present work a CFD solver was developed and named HoAc (an abbreviation<br />

for High order Accuracy) based on the DG discretization. The code was<br />

verified and its ability to achieve the design order <strong>of</strong> the DG discretization for up<br />

to fifth order accurate expansion <strong>of</strong> the solution for the Euler and Navier-Stokes<br />

equations was demonstrated. Furthermore, in the examined cases the ability <strong>of</strong> the<br />

HoAc solver to provide very accurate solutions on problems with complex flow<br />

features was demonstrated.<br />

The HoAc solver has the ability to perform parallel computations via the<br />

domain decomposition method on super computers and massively parallel systems<br />

using the MPI programming paradigm. It can handle arbitrary mixed-type element<br />

meshes over arbitrary geometries in two dimensions and solve unsteady problems<br />

for the Euler and Navier-Stokes equations with the aid <strong>of</strong> high order explicit and<br />

implicit time marching schemes. This was mainly achieved by the fact that all the<br />

operations for the DG discretization in HoAc are being performed in the standard<br />

square element configuration where the expansion bases are defined and constructed.<br />

Also, the capability to produce a high order discontinuous output <strong>of</strong> the solution has<br />

been implemented, which appeared to be fairly adequate for coarse meshes when a


CHAPTER 8. CONCLUSIONS AND FUTURE WORK 199<br />

high order approximation <strong>of</strong> the solution is employed.<br />

Furthermore, in the present work the shock capturing capability <strong>of</strong> the DG<br />

discretization was addressed and improved with a new limiting approach which<br />

is suitable for mixed-type meshes. A shock capturing p-adaptive DG scheme was<br />

developed and implemented in HoAc. In the new limiting approach, limiting is performed<br />

with a slight but essential modification <strong>of</strong> a TVB limiter extensively tested<br />

for DG discretizations. Limiting is carried out in the standard computational space<br />

where all types <strong>of</strong> elements transform to standard squares. Application <strong>of</strong> the new<br />

limiting approach for standard test cases for shock capturing schemes demonstrated<br />

that the new approach is robust, accurate, and it does not require adjustment <strong>of</strong><br />

parameters. Moreover, it provides capabilities for p-adaptation because the solution<br />

accuracy could be increased away from regions <strong>of</strong> discontinuities that are clearly<br />

marked since accurate estimates <strong>of</strong> the second derivative are obtained. In addition,<br />

the developed limiting algorithm was combined with new positivity preserving limiters<br />

especially developed for the DG method and test cases involving high speed<br />

flows with strong shocks were examined with exceptional results. Specifically, these<br />

cases were solved using unstructured triangular meshes and up to the time <strong>of</strong> the<br />

present work nowhere in the literature the positivity preserving limiters appeared<br />

to be applied on triangular meshes.<br />

Future work in the DG discretization and implementation <strong>of</strong> the current thesis<br />

will involve: (i) the extension <strong>of</strong> the developed limiting approach in three dimensions,<br />

both for shock capturing and positivity preservation for density and pressure, (ii)<br />

the ability <strong>of</strong> the HoAc solver to work on meshes with curved elements, (iii) the<br />

extension <strong>of</strong> the HoAc solver to handle three dimensional flows around complex<br />

geometries using mixed-type element meshes with curved elements, (iv) the ability<br />

to perform high order DG simulations for turbulent flows and (v) the extension<br />

<strong>of</strong> the current implementation <strong>of</strong> HoAc for its use with Graphics Processor Units<br />

(GPUs).


APPENDIX A<br />

Mesh quality index Q x for the stretched unstructured mesh <strong>of</strong> Fig. 3.5(b):<br />

Q x = 96<br />

7<br />

dx<br />

∣|1 + 2dx| (1 + 2dx) + 4(1 + dx 2 ) + |2dx − 1| (1 − 2dx) ∣ +<br />

∣ 32√ ∣∣∣<br />

dy<br />

3<br />

7 (2dy − √ 3)[|2dy − √ 3| + | √ 3 + 2dy|] + 2dy<br />

∣ .<br />

(A-1)<br />

Mesh quality index Q x for the skewed mesh <strong>of</strong> Fig. 3.6:<br />

Q x = 1 − cos2 (ω)<br />

2[cos 2 (ω) + 1] + cos (ω) − 1<br />

cos 2 (ω) − cos (ω) − 4 −<br />

3 sin (2ω)<br />

2[cos (ω) + 1][cos 2 (ω) − cos (ω) − 4] .<br />

(A-2)<br />

200


APPENDIX B<br />

Discontinuous Output for Quadrilateral Elements<br />

For an approximation order N, a quadrilateral element is subdivided into (N + 1) 2<br />

quadrilateral sub-elements. This element division is based on a listing Q q <strong>of</strong> equally<br />

spaced points according to Eq. (B-1):<br />

Q q = p + q(N + 2), 0 ≤ p, q ≤ N + 1, (B-1)<br />

over the standard square element configuration. Each point then defines a node<br />

whose coordinates in the physical space are computed by Eq. (4.29). Then the<br />

connectivity <strong>of</strong> the quadrilateral sub-elements is defined according to a structured<br />

mesh. Finally, the solution is projected according to Eq. (4.5) and printed at every<br />

new node. This leads to a print out <strong>of</strong> the solution that is discontinuous across the<br />

inter-element boundaries.<br />

Discontinuous Output for Triangular elements<br />

As for a quadrilateral element a triangular element is subdivided into (N + 1) 2<br />

triangular sub-elements whose nodes are based on a listing Q t <strong>of</strong> equally spaced<br />

points over the standard triangular element configuration given in the following<br />

equation:<br />

201


CHAPTER 8. CONCLUSIONS AND FUTURE WORK 202<br />

Q t = p + (N + 2)q −<br />

q(q − 1)<br />

, 0 ≤ p, q ≤ N with p + q ≤ N. (B-2)<br />

2<br />

The physical coordinates <strong>of</strong> the nodes are computed from Eq. (4.31) and the following<br />

pseudo code is provided for the connectivity list <strong>of</strong> the triangular sub-elements:<br />

m=0;<br />

for (i=0; i< # elements; i++) { // loop elements<br />

for (q = 0; q < N+1; q++)<br />

for (p = 0; p < N+1; p++)<br />

if (p + q < N + 1) { // if in tria<br />

print(Q_t(p, q) + m, Q_t(p+1, q, sys) + m, Q_t(p,q + 1, sys) + m);<br />

if (p + q + 2 < N + 2)<br />

print(Q_t(p + 1, q) + m, Q_t(p + 1, q + 1) + m, Q_t(p, q + 1) + m);<br />

} // if in tria<br />

m += (N + 2) * (N + 3) / 2;<br />

} // loop elments<br />

As for the quadrilateral elements, the solution is projected according to Eq. (4.5)<br />

and printed at every new node resulting in a discontinuous solution print.


APPENDIX C<br />

Convergence analysis<br />

The discretization error is defined as:<br />

Du = u n − u = O (h p ) ,<br />

(C-1)<br />

where p is the order <strong>of</strong> the discretization and u n and u is the numerical solution<br />

obtained with a discretization size h and the exact solution respectively.<br />

The exact solution is necessary in order to compute the order <strong>of</strong> the discretization,<br />

but this is limited to cases that can be solved analytically. The formal order p may<br />

be determined by comparing two discretization errors with differenet sizes h 1 and<br />

h 2 , as follows:<br />

D 1<br />

which after applying logarithms yields:<br />

= hp 1<br />

, (C-2)<br />

D 2<br />

h p 2<br />

and p may be calculated as:<br />

ln D 1<br />

= ln hp 1<br />

, (C-3)<br />

D 2<br />

h p 2<br />

p = − ln D 1 − ln D 2<br />

ln h 1 − ln h 2<br />

.<br />

(C-4)<br />

203


CHAPTER 8. CONCLUSIONS AND FUTURE WORK 204<br />

In Eq. (C-4) scalar error values are needed, which are introduced as integral<br />

values <strong>of</strong> a norm ||∆u|| p <strong>of</strong> the solution error over each element:<br />

||∆u|| p = 1 ∑<br />

(∫<br />

N e<br />

n<br />

Ω e<br />

[Du] p dΩ e<br />

) 1/p<br />

, 1 ≤ n ≤ N e . (C-5)<br />

where N e is the total number <strong>of</strong> elements used in the mesh.


APPENDIX D<br />

HoAc run time options<br />

HoAc relies on PETSc for parallel computations and time marching <strong>of</strong> the solution.<br />

So, any run time option referring to PETSc is readily applied to HoAc as well. A<br />

sample input file for HoAc is given below:<br />

*CONTROL_EQUATION<br />

/*[0,1]=[Euler,NS]*/<br />

1<br />

*END<br />

*CONTROL_FLOWCOND<br />

/*gamma*/ /*P_inf*/ /*U_inf*/ /*V_inf*/ /*rho_inf*/ /*C_inf*/<br />

1.400000e+0 7.142857e-01 0.2 0.000000e+00 1.000000e+00<br />

1.000000e+00<br />

*END<br />

*CONTROL_SOLUTION<br />

/*n_out*/ /*PN*/ /*[0,1]=[start,restart]*/ /*[0,1]=[Roe,LLF]*/ /* # parts*/<br />

5000 1 0 1 16<br />

*END<br />

*CONTROL_TIME<br />

/*End time */ /*Problem type */ /* Output type<br />

[0,1]=[continuous,discontinuous]*/<br />

1.0 0 1<br />

*END<br />

205


CHAPTER 8. CONCLUSIONS AND FUTURE WORK 206<br />

*CONTROL_BOUNDARY_WALL<br />

/*[0,1]=[Straight,CBC]*/<br />

0<br />

*END<br />

*CONTROL_NAVIER_STOKES<br />

/*Re*/ /*Pr*/ /*T_inf*/ /*LDG coefficient*/<br />

100.000000 7.200000e-01 1.0 0.5<br />

*END<br />

*CONTROL_BOUNDARY_IN_SPEC<br />

/*rho_spec_in*/ /*u_spec_in*/ /*v_spec_in*/ /*p_spec_in*/<br />

0.000000e+00 0.000000e+00 0.000000e+00 0.000000e+00<br />

*END<br />

*CONTROL_BOUNDARY_OUT_SPEC<br />

/*rho_spec_out*/ /*u_spec_out*/ /*v_spec_out*/ /*p_spec_out*/<br />

0.000000e+00 0.000000e+00 0.000000e+00 0.000000e+00<br />

*END<br />

where form the comments provided in, the user is easy to check what a value means.<br />

Furthermore, in order to give more flexibility to the user additional run time options<br />

are provided which handle the DG discretization in HoAc . These are given in the<br />

following sample options file along with other PETSc commands for an implicit or<br />

explicit run, where it is noted that the comments should not exist when a simulation<br />

is to be performed.<br />

-llf_flux %% for Local Lax Friedrichs<br />

-roe_flux %% for Roe flux<br />

-explicit %% for an expicit run<br />

-explicit_type 0 %% For using the SSP RK develop by Shu and Osher<br />

-explicit_type 1 %% For using the SSP RK provided by PETSc<br />

-gl %% for Gauss-Legendre quadrature<br />

-gll %% for Gauss-Lobatto-Legendre quadrature<br />

-ts_type ssp


CHAPTER 8. CONCLUSIONS AND FUTURE WORK 207<br />

-ts_ssp_type rks3<br />

-ts_ssp_nstages 9<br />

-ts_dt 1.0e-4<br />

-dt 1.0e-4 %% if explicit_type is 0<br />

-order 1 %% Order <strong>of</strong> approximation<br />

-parts 16 %% Number <strong>of</strong> partitions<br />

-end_time 2000.0 %% End time <strong>of</strong> simulation<br />

-n_out 50000 %% Output frequency<br />

-shock_lim %% Limiter for shock capturing<br />

-pos_lim %% Limiters for positivity preservation<br />

-restart %% Switch for restarting a solution<br />

-snes_monitor<br />

-snes_atol 1.0e-8<br />

-snes_rtol 1.0e-10<br />

-snes_mf<br />

-ksp_type fgmres<br />

-pc_type none<br />

-mat_mffd_compute_normu<br />

-ksp_max_it 30<br />

-ksp_gmres_restart 200<br />

HoAc :<br />

At the command line the user then types the following command for running<br />

mpiexec -n {# parts} ./hoac {input file} -options_file {file_name}


Bibliography<br />

[1] R. Abgrall. On Essentially Non-oscillatory Schemes on Unstructured Meshes:<br />

Analysis and Implementation. Journal <strong>of</strong> Computational Physics, 114(1):45–<br />

58, 1994.<br />

[2] N. Adams. Direct simulation <strong>of</strong> the turbulent boundary layer along a compression<br />

ramp at M= 3 and Re θ = 1685. Journal <strong>of</strong> Fluid Mechanics, 420:47–83,<br />

2000.<br />

[3] M. Ainsworth and J. T. Oden. A Posteriori Error Estimation in Finite Element<br />

Analysis. Wiley-Interscience, NY, 2000.<br />

[4] J. D. Anderson. Computational Fluid Dynamics. The Basics with Applications.<br />

Aeronautical and Aerospace Engineering. McGraw-Hill Inc., 1995.<br />

[5] G. Arabatzis, P. Vavilis, I. Toulopoulos, and J. A. Ekaterinaris. Implicit highorder<br />

time marching schemes for the linearized Euler equations. AIAA Journal,<br />

45(8):1819–1826, 2007.<br />

[6] D. N. Arnold, F. Brezzi, B. Cockburn, and L. D. Marini. Unified Analysis<br />

<strong>of</strong> Discontinuous Galerkin Methods for Elliptic Problems. SIAM J. Numer.<br />

Anal., 39(5):1749–1779, May 2001.<br />

[7] W. E. Arnoldi. The principle <strong>of</strong> minimized iterations in the solution <strong>of</strong> the<br />

matrix eigenvalue problem. Quart. Appl. Math, 9(1):17–29, 1951.<br />

[8] B. Ataie-Ashtiani and S. Hosseini. Error analysis <strong>of</strong> finite difference meth-<br />

208


BIBLIOGRAPHY 209<br />

ods for two-dimensional advection-dispersion-reaction equation. Advances in<br />

Water Resources, 28(8):793–806, 2005.<br />

[9] H. L. Atkins. Continued development <strong>of</strong> the discontinuous Galerkin method<br />

for computational aeroacoustic applications. AIAA paper, 1581:12–14, 1997.<br />

[10] I. Babuska and W. C. Rheinboldt. Error estimates for adaptive finite element<br />

computations. SIAM Journal <strong>of</strong> Numerical Analysis, 15(4):736–754, 1978.<br />

[11] T. J. Baker. Adaptive modification <strong>of</strong> time evolving meshes. Computer Methods<br />

in Applied Mechanics and Engineering, 194(48-49):4977–5001, 2005. Unstructured<br />

Mesh Generation.<br />

[12] D. Balsara and C. Shu. Monotonicity preserving weighted essentially nonoscillatory<br />

schemes with increasingly high order <strong>of</strong> accuracy. Journal <strong>of</strong> Computational<br />

Physics, 160(2):405–452, 2000.<br />

[13] R. Barrett, M. Berry, T. F. Chan, J. Demmel, J. Donato, J. Dongarra, V. Eijkhout,<br />

R. Pozo, C. Romine, and H. V. der Vorst. Templates for the Solution<br />

<strong>of</strong> Linear Systems: Building Blocks for Iterative Methods. SIAM, Philadelphia,<br />

2 edition, 1994.<br />

[14] T. Barth and P. O. Frederickson. High Order Solution <strong>of</strong> the Euler Equations<br />

on Unstructured grids using Quadratic Reconstruction. AIAA Paper 90-0013,<br />

1990.<br />

[15] T. J. Barth. Aspects <strong>of</strong> Unstructured Grids and Finite Volume Solvers for the<br />

Euler and Navier–Stokes equations. In VKI Lecture Series, 1995.<br />

[16] F. Bassi and S. Rebay. Accurate 2D Euler computations by means <strong>of</strong> a high<br />

order discontinuous finite element method. In S. Deshpande, S. Desai, and<br />

R. Narasimha, editors, Fourteenth International Conference on Numerical<br />

Methods in Fluid Dynamics, volume 453 <strong>of</strong> Lecture Notes in Physics, pages<br />

234–240. Springer Berlin / Heidelberg, 1995.


BIBLIOGRAPHY 210<br />

[17] F. Bassi and S. Rebay. A high-order accurate discontinuous finite element<br />

method for the numerical solution <strong>of</strong> the compressible Navier–Stokes equations.<br />

Journal <strong>of</strong> Computational Physics, 131(2):267–279, 1997.<br />

[18] F. Bassi and S. Rebay. High-order accurate discontinuous finite element solution<br />

<strong>of</strong> the 2D Euler equations. Journal <strong>of</strong> Computational Physics, 138(2):251–<br />

285, 1997.<br />

[19] F. Bassi and S. Rebay. An implicit high-order discontinuous Galerkin method<br />

for the steady state compressible Navier-Stokes equations. In Computational<br />

fluid dynamics’98, Proceedings <strong>of</strong> the Fourth European Computational Fluid<br />

Dynamics Conference, pages 1226–1233. Wiley, New York, 1998.<br />

[20] F. Bassi and S. Rebay. Numerical evaluation <strong>of</strong> two discontinuous Galerkin<br />

methods for the compressible Navier–Stokes equations. International journal<br />

for numerical methods in fluids, 40(1-2):197–207, 2002.<br />

[21] K. Bathe. Finite element procedures, volume 2. Prentice hall Englewood Cliffs,<br />

NJ, 1996.<br />

[22] C. E. Baumann and J. T. Oden. A discontinuous hp finite element method<br />

for convection–diffusion problems. Computer Methods in Applied Mechanics<br />

and Engineering, 175(3-4):311–341, July 1999.<br />

[23] C. E. Baumann and J. T. Oden. A discontinuous hp finite element method for<br />

the Euler and Navier–Stokes equations. International journal for numerical<br />

methods in fluids, 31(1):79–95, 1999.<br />

[24] T. Belytschko, W. K. Liu, and B. Moran. Nonlinear finite elements for continua<br />

and structures, volume 36. Wiley, 2000.<br />

[25] M. J. Berger and P. Collela. An Adaptive Multigrid Method for the Euler<br />

Equations. Lecture Notes in Physics, 218:92–97, 1985.<br />

[26] M. J. Berger and P. Collela. Local Adaptive Mesh Refinement for Shock<br />

Hydrodynamics. Journal <strong>of</strong> Computational Physics, 82:64–84, 1989.


BIBLIOGRAPHY 211<br />

[27] C. Bernardo and C.-W. Shu. P 1 RKDG method for two-dimensional Euler<br />

equations <strong>of</strong> gas dynamics. In Fouth International Symposium on CFD, UC<br />

Davis, 1991.<br />

[28] C. Bernardo and C.-W. Shu. The Runge-Kutta Discontinuous Galerkin<br />

Method for Conservation Laws V: Multidimensional Systems. Journal <strong>of</strong> Computational<br />

Physics, 141(2):199–224, 1998.<br />

[29] R. Biswas, K. D. Devine, and J. E. Flaherty. Parallel, adaptive finite element<br />

methods for conservation laws. Applied Numerical Mathematics, 14(1-3):255–<br />

283, 1994.<br />

[30] P. N. Brown and A. C. Hindmarsh. Matrix-free methods for stiff systems <strong>of</strong><br />

ODE’s. SIAM J. Numer. Anal., 23(3):610–638, June 1986.<br />

[31] P. N. Brown and Y. Saad. Hybrid Krylov methods for nonlinear systems <strong>of</strong><br />

equations. SIAM J. Sci. Stat. Comput., 11(3):450–481, May 1990.<br />

[32] A. Burbeau, P. Sagaut, and C. H. Bruneau. A Problem-Independent Limiter<br />

for High-Order Runge-Kutta Discontinuous Galerkin Methods. Journal <strong>of</strong><br />

Computational Physics, 169(1):111–150, 2001.<br />

[33] T. F. Chan and K. R. Jackson. Nonlinearly preconditioned Krylov subspace<br />

methods for discrete Newton algorithms. SIAM Journal on Scientific and<br />

Statistical Computing, 5:533, 1984.<br />

[34] S. Chang and K. Chang. On the shock–vortex interaction in Schardin’s problem.<br />

Shock Waves, 10(5):333–343, 2000.<br />

[35] A. Chaudhuri, A. Hadjadj, and A. Chinnayya. On the use <strong>of</strong> immersed boundary<br />

methods for shock/obstacle interactions. J. Comput. Physics, 230(5):1731–<br />

1748, 2011.<br />

[36] T. Chisholm and D. Zingg. A Jacobian-free Newton-Krylov algorithm for compressible<br />

turbulent fluid flows. Journal <strong>of</strong> Computational Physics, 228(9):3490–<br />

3507, 2009.


BIBLIOGRAPHY 212<br />

[37] P. G. Ciarlet. The Finite Element Method for Elliptic Problems. SIAM, 40,<br />

2002.<br />

[38] B. Cockburn, B. Dong, J. Guzmán, M. Restelli, and R. Sacco. A hybridizable<br />

discontinuous Galerkin method for steady-state convection-diffusion-reaction<br />

problems. SIAM Journal on Scientific Computing, 31:3827, 2009.<br />

[39] B. Cockburn, J. Gopalakrishnan, and R. Lazarov. Unified hybridization <strong>of</strong><br />

discontinuous Galerkin, mixed, and continuous Galerkin methods for second<br />

order elliptic problems. SIAM J. Numer. Anal, 47(2):1319–1365, 2009.<br />

[40] B. Cockburn, S.-Y. Lin, and C.-W. Shu. TVB Runge-Kutta local projection<br />

discontinuous Galerkin finite element method for conservation laws III: Onedimensional<br />

systems. Journal <strong>of</strong> Computational Physics, 84(1):90–113, 1989.<br />

[41] B. Cockburn and C.-W. Shu. TVB Runge-Kutta Local Projection Discontinuous<br />

Galerkin Finite Element Method for Conservation Laws II: General<br />

framework. Mathematics <strong>of</strong> Computation, 52:411–435, 1989.<br />

[42] B. Cockburn and C.-W. Shu. The Runge-Kutta local projection P 1 -<br />

discontinous Galerkin method for scalar conservation laws. M 2 AN, 25:337–<br />

361, 1991.<br />

[43] B. Cockburn and C.-W. Shu. The Local Discontinuous Galerkin Method<br />

for Time-Dependent Convection-Diffusion Systems. SIAM J. Numer. Anal.,<br />

35(6):2440–2463, October 1998.<br />

[44] B. Cockburn, C.-W. Shu, and S. Hou. The Runge-Kutta local projection<br />

discontinuous Galerkin finite element method for conservation laws IV: The<br />

multi-dimensional case. Mathematics <strong>of</strong> Computation, 54:545–581, 1990.<br />

[45] S. Collis. Discontinuous Galerkin methods for turbulence simulation. In Proceedings<br />

<strong>of</strong> the 2002 Center for Turbulence Research Summer Program, pages<br />

155–167, 2002.


BIBLIOGRAPHY 213<br />

[46] A. Crivellini and F. Bassi. An implicit matrix-free Discontinuous Galerkin<br />

solver for viscous and turbulent aerodynamic simulations. Computers & Fluids,<br />

50:81–93, 2011.<br />

[47] M. Delanaye and Y. Liu. Quadratic reconstruction finite volume schemes on<br />

3D arbitrary unstructured polyhedral grids. In Collection <strong>of</strong> Technical Papers.<br />

Vol. 1 (A99-33467 08-34), number AIAA-1999-776 in Computational Fluid<br />

Dynamics Conference, 14th, Norfolk, VA, June 28-July 1 1999.<br />

[48] J. Dompierre, M. G. Vallet, P. Labbé, and F. Guibalt. An analysis <strong>of</strong> simplex<br />

shape measures for anisotropic meshes. Computer Methods in Applied<br />

Mechanics and Engineering, Vol. 194:pp. 4895–4914, 2005.<br />

[49] M. Dubiner. Spectral Methods on Triangles and Other Domains. Journnal <strong>of</strong><br />

Scientific Computing, 6(4), 1991.<br />

[50] E. Elsen, P. LeGresley, and E. Darve. Large calculation <strong>of</strong> the flow over<br />

a hypersonic vehicle using a GPU. Journal <strong>of</strong> Computational Physics,<br />

227(24):10148–10161, 2008.<br />

[51] A. Emery. An evaluation <strong>of</strong> several differencing methods for inviscid fluid flow<br />

problems. Journal <strong>of</strong> Computational Physics, 2(3):306–331, 1968.<br />

[52] J. M. Escobar, R. Montenegro, G. Montero, E. Rodriguez, and J. M. Gonzalez-<br />

Yuste. Smoothing and local refinement techniques for improving tetrahedral<br />

mesh quality. Computers and Structures, Vol. 83:pp. 2423–2430, 2005.<br />

[53] Q. H. Fang, M. Y. Hussaini, and P. Rasetarinera. An Analysis <strong>of</strong> the Discontinuous<br />

Galerkin Method for Wave Propagation Problems. Journal <strong>of</strong> Computational<br />

Physics, 151(2):921–946, 1999.<br />

[54] D. V. Gaitonde and M. R. Visbal. Pade-type high-order boundary filters for<br />

the Navier-Stokes equations. AIAA Journal, 38(11):2103–2112, 2000.<br />

[55] R. V. Garimella, M. J. Shashkov, and P. M. Knupp. Triangular and quadrilateral<br />

surface mesh quality optimization using local parametrization. Computer<br />

Methods in Applied Mechanics and Engineering, Vol. 193:pp. 913–928, 2004.


BIBLIOGRAPHY 214<br />

[56] C. W. Gear and Y. Saad. Iterative solution <strong>of</strong> linear equations in ODE codes.<br />

SIAM journal on scientific and statistical computing, 4:583, 1983.<br />

[57] M. B. Giles. Accuracy <strong>of</strong> Node–Based Solutions on Irregular Meshes. In<br />

D. L. Dwoyer, M. Y. Hussaini, and R. G. Voigt, editors, 11th International<br />

Conference on Numerical Methods in Fluid Dynamics. Editor: D. L. Dwoyer,<br />

volume 323 <strong>of</strong> Lecture Notes in Physics, pages 273–277, 1989.<br />

[58] W. D. Gropp, L. C. McInnes, B. F. Smith, and S. Balay. PETSc home page.<br />

http://www.mcs.anl.gov/petsc, 2008.<br />

[59] W. D. Gropp, L. C. McInnes, B. F. Smith, and S. Balay. PETSc Users Manual.<br />

Technical Report ANL-95/11 - Revision 3.2, Argonne National Laboratory,<br />

2011.<br />

[60] S. H. High frequency cinematographhy in the shock tube. Journal <strong>of</strong> Photo<br />

Science, 5:19–26, 1957.<br />

[61] H.C. and Yee. Explicit and Implicit Multidimensional Compact High-<br />

Resolution Shock-Capturing Methods:Formulation. Journal <strong>of</strong> Computational<br />

Physics, 131(1):216–232, 1997.<br />

[62] J. C. Heinrich and O. C. Zienkiewicz. Quadratic finite element schemes for<br />

two-dimensional convective-transport problems. International Journal for Numerical<br />

Methods in Engineering, 11(12):1831–1844, 1977.<br />

[63] C. Hirsch. Numerical Computation <strong>of</strong> Internal and External Flows, Vol. 1<br />

Fundamentals <strong>of</strong> Numerical Discretization, 1988.<br />

[64] C. Hirsch. Numerical computation <strong>of</strong> internal and external flows, Vol. 2, Computational<br />

methods for inviscid and viscous flows, 1990.<br />

[65] T. Hughes, L. Franca, and M. Balestra. A new finite element formulation for<br />

computational fluid dynamics: V. Circumventing the Babuska-Brezzi condition:<br />

A stable Petrov-Galerkin formulation <strong>of</strong> the Stokes problem accommodating<br />

equal-order interpolations. Computer Methods in Applied Mechanics<br />

and Engineering, 59(1):85–99, 1986.


BIBLIOGRAPHY 215<br />

[66] T. Ikeda and P. Durbin. Mesh stretch effects on convection in flow simulations.<br />

Journal <strong>of</strong> Computational Physics, Vol. 199:pp. 110–125, 2004.<br />

[67] C. Ilinca. A comparison <strong>of</strong> three Error Estimation Techniques for Finite–<br />

Volume Solutions <strong>of</strong> compressible Flows. Computer Methods in Applied Mechanics<br />

and Engineering, Vol. 189:pp. 94–121, 2000.<br />

[68] J. Jaffre, C. Johnson, and A. Szepessy. Convergence <strong>of</strong> the discontinuous<br />

Galerkin finite element method for hyperbolic conservation laws. Mathematical<br />

Models and Methods in Applied Sciences, 5(3):367–386, 1995.<br />

[69] A. Jameson. Essential elements <strong>of</strong> computational algorithms for aerodynamic<br />

analysis and design. Technical Report NASA/CR-97206268, ICASE 97-68,<br />

1997.<br />

[70] H. Jasak and A. Gosman. Automatic Resolution Control for the Finite–<br />

Volume Method, Parts 1,2, and 3. Numerical Heat Transfer, Vol. 38:pp.<br />

237–290, 2000.<br />

[71] H. Jasak and A. Gosman. Element Residual Error Estimate for the Finite–<br />

Volume Method. Computers and Fluids, Vol. 32:pp. 223–248, 2003.<br />

[72] Y. N. Jeng and J. L. Chen. Truncation error analysis <strong>of</strong> the finite volume<br />

method for a model steady convective equation. Journal <strong>of</strong> Computational<br />

Physics, 100(1):64–76, 1992.<br />

[73] G.-S. Jiang and C.-W. Shu. Efficient Implementation <strong>of</strong> Weighted ENO<br />

Schemes. Journal <strong>of</strong> Computational Physics, 126(1):202–228, 1996.<br />

[74] J. E. D. Jr. and R. B. Schnabel. Numerical methods for unconstrained optimization<br />

and nonlinear equations. Prentice Hall series in computational mathematics.<br />

Prentice Hall, 1983.<br />

[75] Y. Kallinderis. Numerical Treatment <strong>of</strong> Grid Interfaces for Viscous Flows.<br />

Journal <strong>of</strong> Computational Physics, Vol. 98:pp. 129–144, 1992.


BIBLIOGRAPHY 216<br />

[76] Y. Kallinderis. Grid adaptation by redistribution and local embedding. In<br />

VKI Lecture Series, 1996.<br />

[77] Y. Kallinderis. Hybrid Grids and their Applications, chapter 35. Handbook <strong>of</strong><br />

Grid Generation. CRC Press, Boca Raton, FL, 1999.<br />

[78] Y. Kallinderis and J. R. Baron. A new adaptive algorithm for turbulent flows.<br />

Computers & Fluids, 21(1):77–96, 1992.<br />

[79] Y. Kallinderis and C. Kontzialis. A priori mesh quality estimation via direct<br />

relation between truncation error and mesh distortion. Journal <strong>of</strong> Computational<br />

Physics, 228(3):881–902, 2009.<br />

[80] Y. Kallinderis and S. Kwong. Viscous Grids and Assesment <strong>of</strong> their Quality.<br />

In 15 th AIAA Computational Fluid Dynamics Conference, AIAA–2001–2539,<br />

June 2001.<br />

[81] K. Karantasis, E. Polychronopoulos, K. Panourgias, and J. Ekaterinaris. Accelerating<br />

the simulation <strong>of</strong> brain tumor proliferation with many-core GPUs.<br />

Journal <strong>of</strong> Computational Science, 2011.<br />

[82] G. E. Karniadakis, B. Cockburn, and C. W. Shu. The development <strong>of</strong> discontinuous<br />

Galerkin method. Number 11 in Lecture Notes in Computational<br />

Science and Engineering. Springer, 2000.<br />

[83] G. E. Karniadakis and S. J. Sherwin. Spectral/hp Element Methods for CFD.<br />

Oxford <strong>University</strong> Press, 2nd edition.<br />

[84] C. T. Kelley. Iterative Methods for Linear and Nonlinear Equations. Number<br />

16 in Frontiers in Applied Mathematics. SIAM, 1995.<br />

[85] D. Ketcheson and A. Robinson. On the practical importance <strong>of</strong> the SSP<br />

property for Runge–Kutta time integrators for some common Godunov-type<br />

schemes. International Journal for Numerical Methods in Fluids, 48(3):271–<br />

303, 2005.


BIBLIOGRAPHY 217<br />

[86] C. M. Klaij, J. J. W. Van der Vegt, and H. Van der Ven. Space-time discontinuous<br />

Galerkin method for the compressible Navier-Stokes equations. Journal<br />

<strong>of</strong> Computational Physics, 217(2):589–611, 2006.<br />

[87] A. Klöckner, T. Warburton, J. Bridge, and J. Hesthaven. Nodal discontinuous<br />

Galerkin methods on graphics processors. Journal <strong>of</strong> Computational Physics,<br />

228(21):7863–7882, 2009.<br />

[88] D. Knoll and D. Keyes. Jacobian-free Newton–Krylov methods: a survey <strong>of</strong><br />

approaches and applications. Journal <strong>of</strong> Computational Physics, 193(2):357–<br />

397, 2004.<br />

[89] K. Kontzialis and J. A. Ekaterinaris. Unsteasy viscous flow over tandem<br />

NACA0012 airfoils with a smooth initial condition. In 1st International Workshop<br />

on High Order CFD Methods, Nashville, Tennessee, 2012.<br />

[90] K. Kontzialis, K. Panourgias, and J. Ekaterinaris. A Limiting Approach for<br />

Three-Dimensional DG Discetizations in Arbitrary-Type Meshes . In 20th<br />

AIAA Computational Fluid Dynamics Conference, Honolulu, Hawaii, number<br />

AIAA-2011-3837, 2011.<br />

[91] L. Krivodonova. Limiters for high-order discontinuous Galerkin methods.<br />

Journal <strong>of</strong> Computational Physics, 226(1):879–896, 2007.<br />

[92] L. Krivodonova and M. J. Berger. High-order accurate implementation <strong>of</strong> solid<br />

wall boundary conditions in curved geometries. Journal <strong>of</strong> Computational<br />

Physics, 211(2):492–512, 2006.<br />

[93] B. Landmann. A parallel discontinuous Galerkin code for the Navier-Stokes<br />

and Reynoldsaveraged Navier-Stokes equations. PhD thesis, <strong>University</strong> <strong>of</strong> Stutgard,<br />

2008.<br />

[94] A. R. Larzelere. Creating simulation capabilities. Computational Science &<br />

Engineering, IEEE, 5(1):27–35, 1998.<br />

[95] S. Lele. Compact finite difference schemes with spectral-like resolution. Journal<br />

<strong>of</strong> Computational Physics, 103(1):16–42, 1992.


BIBLIOGRAPHY 218<br />

[96] R. J. Leveque. Finite Volume Methods for Hyperbolic Problems. Cambridge<br />

Texts in Applied Mathematics. Cambride <strong>University</strong> Press, 2002.<br />

[97] R. J. LeVeque. Finite difference methods for ordinary and partial differential<br />

equations - steady-state and time-dependent problems. SIAM, 2007.<br />

[98] Y. Liu, M. Vinokur, and Z. J. Wang. Spectral difference method for unstructured<br />

grids I: Basic formulation. Journal <strong>of</strong> Computational Physics,<br />

216(2):780–801, 2006.<br />

[99] W. Lowrie, V. S. Lukin, and U. Shumlak. A priori mesh quality metric error<br />

analysis applied to a high-order finite element method. Journal <strong>of</strong> Computational<br />

Physics, 230(14):5564–5586, 2011.<br />

[100] H. Luo, J. D. Baum, and R. Löhner. A Hermite WENO-based limiter for discontinuous<br />

Galerkin method on unstructured grids. Journal <strong>of</strong> Computational<br />

Physics, 225(1):686–713, 2007.<br />

[101] Matlab. Symbolic Math Toolbox. The Matlab Inc., 1998.<br />

[102] D. McRae. r-Refinement grid adaptation algorithms and issues. Computer<br />

Methods in Applied Mechanics and Engineering, Vol. 189:pp. 1161–1182, 2000.<br />

[103] C. R. Nastase and D. J. Mavriplis. High-order discontinuous Galerkin methods<br />

using an hp-multigrid approach. Journal <strong>of</strong> Computational Physics,<br />

213(1):330–357, 2006.<br />

[104] N. C. Nguyen, J. Peraire, and B. Cockburn. An implicit high-order hybridizable<br />

discontinuous Galerkin method for nonlinear convection-diffusion equations.<br />

Journal <strong>of</strong> Computational Physics, 228(23):8841–8855, 2009.<br />

[105] N. C. Nguygen, P. Persson, and J. Peraire. RANS solutions using high order<br />

discontinous Galerkin methods. AIAA-2007-914, 2007.<br />

[106] J. T. Oden, I. Babuska, F. Nobile, Y. Feng, and R. Tempone. Theory and<br />

methodology for estimation and control <strong>of</strong> errors due to modeling, approximation,<br />

and uncertainty. Computer Methods in Applied Mechanics and Engi-


BIBLIOGRAPHY 219<br />

neering, 194(2-5):195–204, 2005. Selected papers from the 11th Conference on<br />

The Mathematics <strong>of</strong> Finite Elements and Applications.<br />

[107] J. T. Oden, I. Babuka, and C. E. Baumann. A DiscontinuoushpFinite Element<br />

Method for Diffusion Problems. Journal <strong>of</strong> Computational Physics,<br />

146(2):491–519, 1998.<br />

[108] J. T. Oden, W. Wu, and M. Ainsworth. An a posteriori error estimate for finite<br />

element approximations <strong>of</strong> the Navier-Stokes equations. Computer Methods<br />

in Applied Mechanics and Engineering, 111(1-2):185–202, 1994.<br />

[109] U. D. <strong>of</strong> Energy. Scientific discovery through advanced computing. Technical<br />

report, U.S. Deprtment <strong>of</strong> Energy, 2000.<br />

[110] F. Oliver. Weighted Essentially Non-Oscillatory Schemes for the Interpolation<br />

<strong>of</strong> Mean Values on Unstructured Grids. Journal <strong>of</strong> Computational Physics,<br />

144(1):194–212, 1998.<br />

[111] M. Patrick and Knupp. Algebraic mesh quality metrics for unstructured initial<br />

meshes. Finite Elements in Analysis and Design, Vol. 39:pp. 217–241, 2003.<br />

[112] B. Perthame and C. W. Shu. On positivity preserving finite volume schemes<br />

for Euler equations. Numerische Mathematik, 73(1):119–130, 1996.<br />

[113] T. J. Poinsot and S. K. Lele. Boundary conditions for direct simulations <strong>of</strong><br />

compressible viscous flows. Journal <strong>of</strong> computational physics, 101(1):104–129,<br />

1992.<br />

[114] M. Proot. The least-squares spectral element method. PhD thesis, PhD thesis,<br />

Delft <strong>University</strong> <strong>of</strong> Technology, 2003.<br />

[115] M. Proot and M. I. Gerrtisma. Least-squares spectral elements applied to the<br />

Stokes problem. Journal <strong>of</strong> Computational Physics, 181(2):454–477, 2002.<br />

[116] S. Prudhomme and J. T. Oden. On goal-oriented error estimation for elliptic<br />

problems: application to the control <strong>of</strong> pointwise errors. Computer Methods<br />

in Applied Mechanics and Engineering, Vol. 176:pp. 313–331, 1999.


BIBLIOGRAPHY 220<br />

[117] S. Prudhomme and J. T. Oden. A posteriori Error Estimation in Fluid Mechanics.<br />

In Proceedings <strong>of</strong> the Finite Elements in Flow Problems 2000 Conference,<br />

The <strong>University</strong> <strong>of</strong> Texas at Austin, April 30 - May 4, 2000.<br />

[118] T. H. Pulliam. Solution methods in computational fluid dynamics. Notes<br />

for the von Kármán Institute For Fluid Dynamics Lecture Series: Numerical<br />

Techniques for Viscous Flow Computation In Turbomachinery Blades, 1985.<br />

[119] J. Qiu, B. C. Khoo, and C.-W. Shu. A numerical study for the performance <strong>of</strong><br />

the Runge–Kutta discontinuous Galerkin method based on different numerical<br />

fluxes. Journal <strong>of</strong> Computational Physics, 212(2):540–565, 2006.<br />

[120] J. Qiu and C.-W. Shu. Hermite WENO schemes and their application as limiters<br />

for Runge-Kutta discontinuous Galerkin method: one-dimensional case.<br />

Journal <strong>of</strong> Computational Physics, 193(1):115–135, 2004.<br />

[121] P. Rasetarinera and M. Y. Hussaini. An Efficient Implicit Discontinuous Spectral<br />

Galerkin Method. Journal <strong>of</strong> Computational Physics, 172(2):718–738,<br />

2001.<br />

[122] W. H. Reed and T. R. Hill. Triangular mesh methods for the neutorn transport<br />

equation. Technical report, Los Alamos Scientific Laboaratory, 1973.<br />

[123] J. K. Reid. On the method <strong>of</strong> conjugate gradients for the solution <strong>of</strong> large<br />

sparse systems <strong>of</strong> linear equations, pages 231–254. Academic Press, 1971.<br />

[124] P. Roe. Approximate Riemann solvers, parameter vectors, and difference<br />

schemes. Journal <strong>of</strong> computational physics, 43(2):357–372, 1981.<br />

[125] Y. Saad. Iterative Methods for Sparse Linear Systems. PWS Publishing Company,<br />

Boston, 1996.<br />

[126] Y. Saad and M. H. Schultz. GMRES: A generalized minimal residual algorithm<br />

for solving nonsymmetric linear systems. SIAM J. Sci. Stat. Comput.,<br />

7(3):856–869, 1986.


BIBLIOGRAPHY 221<br />

[127] S. Sankaranarayanan and M. L. Spaulding. A Study <strong>of</strong> the Effects <strong>of</strong> Grid<br />

Non–Orthogonality on the Solution <strong>of</strong> Shallow Water Equations in Boundary–<br />

Fitted Coordinate Systems. Journal <strong>of</strong> Computational Physics, Vol. 184:pp.<br />

299–320, 2003.<br />

[128] J. Shi, Y.-T. Zhang, and C.-W. Shu. Resolution <strong>of</strong> high order WENO<br />

schemes for complicated flow structures. Journal <strong>of</strong> Computational Physics,<br />

186(2):690–696, 2003.<br />

[129] C.-W. Shu. TVB uniformly high-order schemes for conservation laws. Mathematics<br />

<strong>of</strong> Computation, 49:105–121, 1987.<br />

[130] C.-W. Shu and S. Osher. Efficient implementation <strong>of</strong> essentially non-oscillatory<br />

shock-capturing schemes. Journal <strong>of</strong> Computational Physics, 77(2):439–471,<br />

1988.<br />

[131] B. Sjögreen and H. C. Yee. Grid convergence <strong>of</strong> high order methods for multiscale<br />

complex unsteady viscous compressible flows. Journal <strong>of</strong> Computational<br />

Physics, 185(1):1–26, 2003.<br />

[132] J. C. Tannehill, D. A. Anderson, and R. H. Pletcher. Computational Fluid<br />

Mechanics and Heat Transfer. Series in Computational and Physical Processes<br />

in Mechanics and Thermal Sciences. Taylor & Francis, 2nd edition, 1997.<br />

[133] J. F. Thompson. Numerical Grid Generation. North-Holland, New York, 1982.<br />

[134] M. C. Thompson. An Adaptive Multigrid Technique for the Incompressible<br />

Navier–Stokes Equations. Journal <strong>of</strong> Computational Physics, Vol. 82:pp. 94–<br />

121, 1992.<br />

[135] E. F. Toro. Riemann solvers and numerical methods for fluid dynamics: a<br />

practical introduction. Springer Verlag, 2009.<br />

[136] E. F. Toro, M. Spruce, and W. Speares. Restoration <strong>of</strong> the contact surface in<br />

the HLL-Riemann solver. Shock waves, 4(1):25–34, 1994.


BIBLIOGRAPHY 222<br />

[137] I. Toulopoulos and J. A. Ekaterinaris. Discontinuous-galerkin discretizations<br />

for viscous flow problems in complex domains. In 43rd AIAA Aerospace Sciences<br />

Meeting and Exhibit, Reno, Nevada, number AIAA-2005-1264, Jan. 10-<br />

13 2005.<br />

[138] I. Toulopoulos and J. A. Ekaterinaris. High Resolution Compressible Flow<br />

Simulations with the Discontinuous Galerkin Methd. In 17th AIAA Computational<br />

Fluid Dynamics Conference, Toronto, Ontario, number AIAA-2005-<br />

5109, June 6-July 9 2005.<br />

[139] I. Toulopoulos and J. A. Ekaterinaris. High-order discontinuous-Galerkin discretizations<br />

for computational aeroacoustics in complex domains. AIAA Journal,<br />

44(3):502–511, 2006.<br />

[140] I. Toulopoulos and J. A. Ekaterinaris. Artificial boundary conditions for<br />

the numerical solution <strong>of</strong> the Euler equations by the discontinuous galerkin<br />

method. J. Comput. Physics, 230(15):5974–5995, 2011.<br />

[141] E. Turkel. Accuracy <strong>of</strong> Schemes With Nonuniform Meshes for Compressible<br />

Fluid Flows. Applied Numerical Mathematics, Vol. 2:pp. 529–550, 1986.<br />

[142] J. J. W. van der Vegt and H. van der Ven. Discontinuous Galerkin Finite<br />

Element Method with Anisotropic Local Grid Refinement for Inviscid Compressible<br />

Flows. Journal <strong>of</strong> Computational Physics, 141(1):46–77, 1998.<br />

[143] H. van der Ven and J. J. W. van der Vegt. Space-time discontinuous Galerkin<br />

finite element method with dynamic grid motion for inviscid compressible<br />

flows: II. Efficient flux quadrature. Computer Methods in Applied Mechanics<br />

and Engineering, 191(41-42):4747–4780, 2002.<br />

[144] D. A. Venditti and D. L. Darm<strong>of</strong>al. Adjoint Error Estimation and Grid Adaptation<br />

for Functional Outputs: Application to Quasi-One-Dimensional Flow.<br />

Journal <strong>of</strong> Computational Physics, 164(1):204–227, 2000.<br />

[145] V. Venkatakrishnan, S. Allmaras, and D. Kamenetskii. Higher Order Shemes<br />

for the compressible Navier-Stokes Equations. AIAA Paper 2003-3987, 2003.


BIBLIOGRAPHY 223<br />

[146] V. Venkatakrishnan and D. J. Mavriplis. Implicit solvers for unstructured<br />

meshes. Journal <strong>of</strong> Computational Physics, 105(1):83–91, March 1993.<br />

[147] M. R. Visbal and D. V. Gaitonde. On the Use <strong>of</strong> Higher-Order Finite-<br />

Difference Schemes on Curvilinear and Deforming Meshes. Journal <strong>of</strong> Computational<br />

Physics, 181(1):155–185, 2002.<br />

[148] J. Vos, A. Rizzi, D. Darracq, and E. Hirschel. Navier–Stokes solvers in European<br />

aircraft design. Progress in Aerospace Sciences, 38(8):601–697, 2002.<br />

[149] Z. J. Wang, L. Zhang, and Y. Liu. Spectral (finite) volume method for conservation<br />

laws on unstructured grids IV: extension to two-dimensional systems.<br />

Journal <strong>of</strong> Computational Physics, 194(2):716–741, 2004.<br />

[150] T. Warburton. Spectral/hp Methods on Polymorphic Multi-Domains: Algorithms<br />

and Applications. PhD thesis, Brown <strong>University</strong>, 1999.<br />

[151] T. C. Warburton, S. J. Sherwin, and G. E. Karniadakis. Basis Functions for<br />

Triangular and Quadrilateral High-Order Elements. SIAM J. Sci. Comput.,<br />

20(5):1671–1695, April 1999.<br />

[152] P. Woodward and P. Colella. The numerical simulation <strong>of</strong> two-dimensional<br />

fluid flow with strong shocks. Journal <strong>of</strong> Computational Physics, 54(1):115–<br />

173, 1984.<br />

[153] H. C. Yee, N. D. Sandham, and M. J. Djomehri. Low-Dissipative High-Order<br />

Shock-Capturing Methods Using Characteristic-Based Filters,. Journal <strong>of</strong><br />

Computational Physics, 150(1):199–238, 1999.<br />

[154] D. You, R. Mittal, M. Wang, and P. Moin. Analysis <strong>of</strong> stability and accuracy <strong>of</strong><br />

finite-difference schemes on a skewed mesh. Journal <strong>of</strong> Computational Physics,<br />

Vol. 213:pp. 184–204, 2006.<br />

[155] M. Zhang and C. W. Shu. An analysis <strong>of</strong> three different formulations for the<br />

discontinous Galerkin method for diffusion equations. Mathematical Models<br />

and Methods in Applied Sciences, 13(3):395–413, 2003.


BIBLIOGRAPHY 224<br />

[156] X. Zhang and C.-W. Shu. On positivity-preserving high order discontinuous<br />

Galerkin schemes for compressible Euler equations on rectangular meshes.<br />

Journal <strong>of</strong> Computational Physics, 229(23):8918–8934, 2010.<br />

[157] J. Zhu and J. Qiu. Local DG method using WENO type limiters for convectiondiffusion<br />

problems. Journal <strong>of</strong> Computational Physics, In Press, Corrected<br />

Pro<strong>of</strong>:–, 2010.<br />

[158] J. Zhu, J. Qiu, C.-W. Shu, and M. Dumbser. Runge-Kutta discontinuous<br />

Galerkin method using WENO limiters II: Unstructured meshes. Journal <strong>of</strong><br />

Computational Physics, 227(9):4330–4353, 2008.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!