You are on page 1of 117

LARGE EDDY SIMULATIONS OF LAMINAR SEPARATION BUBBLE

FLOWS
by
Francois Cadieux

A Dissertation Presented to the


FACULTY OF THE USC GRADUATE SCHOOL
UNIVERSITY OF SOUTHERN CALIFORNIA
In Partial Fulllment of the
Requirements for the Degree
DOCTOR OF PHILOSOPHY
(AEROSPACE ENGINEERING)

May 2015

Copyright 2015

Francois Cadieux

UMI Number: 3704224

All rights reserved


INFORMATION TO ALL USERS
The quality of this reproduction is dependent upon the quality of the copy submitted.
In the unlikely event that the author did not send a complete manuscript
and there are missing pages, these will be noted. Also, if material had to be removed,
a note will indicate the deletion.

UMI 3704224
Published by ProQuest LLC (2015). Copyright in the Dissertation held by the Author.
Microform Edition ProQuest LLC.
All rights reserved. This work is protected against
unauthorized copying under Title 17, United States Code

ProQuest LLC.
789 East Eisenhower Parkway
P.O. Box 1346
Ann Arbor, MI 48106 - 1346

Acknowledgments
A heartfelt thanks to my supervisor Andrzej for his guidance, support and sense
of humour. To Giacomo, thank you for countless in-depth and useful discussions.
I am also grateful to those who shared their code with me without which this
endeavour might have taken yet another year: Tawan, Brian, Tak, and Peter. I
am deeply indebted to Vina, who was always by my side when I needed her most.
The support of my parents Johanne and Yves and my sister Genevieve and their
eagerness for me to join them in their travels kept me sane and focused. Finally, a
special thanks to Dr Philippe Spalart for sharing his DNS data and answering my
many questions. This research was supported by the National Science Foundation
through grant CBET-1233160.

ii

Contents
Acknowledgments

ii

List of Tables

List of Figures

vi

Abstract

viii

1 Introduction
1.1 Laminar Separation Bubble Flows . . . . . . . . . . . . . . . . . . .
1.2 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1
1
6

2 Background
2.1 Governing Equations . . . . . . . . . . . . . . . . . . . . .
2.2 Direct Numerical Simulation . . . . . . . . . . . . . . . . .
2.3 Reynolds-averaged Navier-Stokes . . . . . . . . . . . . . .
2.3.1 RANS for Laminar Separation Bubble Flows . . . .
2.4 Large Eddy Simulation . . . . . . . . . . . . . . . . . . . .
2.4.1 Filtered Navier-Stokes Equations . . . . . . . . . .
2.4.2 Subgrid-scale Models . . . . . . . . . . . . . . . . .
2.4.3 LES Results for Laminar Separation Bubble Flows
2.5 Wall-modeled LES and Hybrid RANS-LES . . . . . . . . .

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

7
. 7
. 8
. 9
. 11
. 14
. 14
. 15
. 21
. 23

3 Methodology
3.1 Approach . . . . . . . . . . . . . . . . . . . .
3.2 Flow Specication . . . . . . . . . . . . . . . .
3.3 Numerical Methods . . . . . . . . . . . . . . .
3.3.1 Center for Turbulence Research Code .
3.3.2 Spectral code in vorticity form . . . . .
3.3.3 Spectral code in skew-symmetric form
3.4 Validation . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

25
25
27
28
30
32
36
38

iii

4 Center for Turbulence Research Results


46
4.1 Numerical Dissipation . . . . . . . . . . . . . . . . . . . . . . . . . 49
4.1.1 Estimating Numerical Dissipation due to Filtering . . . . . . 52
4.1.2 Quantifying Numerical Dissipation . . . . . . . . . . . . . . 54
5 Spectral Results I
59
5.1 LES at 3% of DNS Resolution . . . . . . . . . . . . . . . . . . . . . 61
5.2 LES at 1% of DNS Resolution . . . . . . . . . . . . . . . . . . . . . 63
5.3 Spanwise Structures . . . . . . . . . . . . . . . . . . . . . . . . . . 72
6 Spectral Results II
6.1 LES-V at 3% of DNS Resolution
6.2 LES-V at 1% of DNS Resolution

83
. . . . . . . . . . . . . . . . . . . 83
. . . . . . . . . . . . . . . . . . . 85

7 Conclusions
96
7.1 Summary and Conclusions . . . . . . . . . . . . . . . . . . . . . . . 96
7.2 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
Reference List

101

iv

List of Tables
3.1

Validation using linear stability theory: vorticity spectral code . . .

40

3.2

Validation using linear stability theory: skew-symmetric spectral code 45

4.1

CTR simulation parameters . . . . . . . . . . . . . . . . . . . . . .

46

4.2

CTR numerical dissipation with ltering . . . . . . . . . . . . . . .

57

4.3

CTR numerical dissipation without ltering . . . . . . . . . . . . .

58

5.1

Spectral simulation parameters . . . . . . . . . . . . . . . . . . . . 60

5.2

Spectral 1% LES performance . . . . . . . . . . . . . . . . . . . . .

69

6.1

Vorticity spectral simulation parameters . . . . . . . . . . . . . . .

84

6.2

Spectral 1% LES-V performance . . . . . . . . . . . . . . . . . . . .

90

List of Figures
1.1

Laminar separation bubble ow sketch . . . . . . . . . . . . . . . .

1.2

Laminar separation bubble on an airfoil . . . . . . . . . . . . . . . .

3.1

Cp of laminar separation bubble on airfoil and at plate . . . . . . .

41

3.2

Anatomy of a at plate laminar separation bubble ow . . . . . . .

42

3.3

Visualization of a at plate laminar separation bubble ow . . . . .

43

3.4

Flat plate laminar separation bubble ow computational setup . . .

43

3.5

Numerical dissipation example . . . . . . . . . . . . . . . . . . . . .

44

3.6

Spectral LES boundary conditions . . . . . . . . . . . . . . . . . . . 44

4.1

CTR boundary conditions . . . . . . . . . . . . . . . . . . . . . . .

48

4.2

Visualization of CTR DNS . . . . . . . . . . . . . . . . . . . . . . .

49

4.3

CTR mean velocity contour plot . . . . . . . . . . . . . . . . . . . .

50

4.4

CTR time-averaged Cp and Cf . . . . . . . . . . . . . . . . . . . . .

51

4.5

CTR energy rate-of-change in turbulent region . . . . . . . . . . . .

53

5.1

Spectral 3% LES ow visualization . . . . . . . . . . . . . . . . . .

61

5.2

Spectral 3% LES Cp and Cf . . . . . . . . . . . . . . . . . . . . . .

64

5.3

Spectral 3% LES boundary layer thicknesses . . . . . . . . . . . . .

65

5.4

Spectral 3% LES mean and RMS velocity proles . . . . . . . . . . 66

vi

5.5

Spectral 1% LES Cp and Cf . . . . . . . . . . . . . . . . . . . . . .

70

5.6

Spectral 1% LES subgrid-scale dissipation . . . . . . . . . . . . . .

75

5.7

Spectral 1% LES boundary layer thicknesses . . . . . . . . . . . . .

76

5.8

Spectral 1% LES mean velocity proles in wall units

77

5.9

Spectral 1% LES mean and RMS velocity proles . . . . . . . . . . 78

. . . . . . . .

5.10 Spectral 1% LES maximum RMS velocity . . . . . . . . . . . . . .

79

5.11 Spectral 1% LES velocity auto-correlation functions . . . . . . . . .

80

5.12 Spectral 1% LES velocity auto-correlations for wider domain . . . .

81

5.13 Spectral 3% LES velocity auto-correlation functions . . . . . . . . .

82

6.1

Spectral 3% LES-V Cp and Cf . . . . . . . . . . . . . . . . . . . . .

86

6.2

Spectral 3% LES-V boundary layer thicknesses . . . . . . . . . . . .

87

6.3

Spectral 3% LES-V mean and RMS velocity proles . . . . . . . . .

88

6.4

Spectral 1% LES-V Cp and Cf . . . . . . . . . . . . . . . . . . . . .

91

6.5

Spectral 1% LES-V boundary layer thicknesses . . . . . . . . . . . .

92

6.6

Spectral 1% LES-V mean velocity proles in wall units . . . . . . .

93

6.7

Spectral 1% LES-V mean and RMS velocity proles . . . . . . . . .

94

6.8

Spectral 1% LES-V maximum RMS velocity . . . . . . . . . . . . .

95

vii

Abstract
The ow over blades and airfoils at moderate angles of attack and Reynolds
numbers ranging from 104 to 105 undergoes separation due to the adverse pressure
gradient generated by surface curvature. In many cases, the separated shear layer
then transitions to turbulence and reattaches, closing o a recirculation region the laminar separation bubble. To avoid body-tted mesh generation problems
and numerical issues, an equivalent problem for ow over a at plate is formulated by imposing boundary conditions that lead to a pressure distribution and
Reynolds number that are similar to those on airfoils. Spalart & Strelets (2000)
tested a number of Reynolds-averaged Navier-Stokes (RANS) turbulence models
for a laminar separation bubble ow over a at plate. Although results with
the Spalart-Allmaras turbulence model were encouraging, none of the turbulence
models tested reliably recovered time-averaged direct numerical simulation (DNS)
results. The purpose of this work is to assess whether large eddy simulation (LES)
can more accurately and reliably recover DNS results using drastically reduced
resolution on the order of 1% of DNS resolution which is commonly achievable for LES of turbulent channel ows. LES of a laminar separation bubble ow
over a at plate are performed using a compressible sixth-order nite-dierence
code and two incompressible pseudo-spectral Navier-Stokes solvers at resolutions
corresponding to approximately 3% and 1% of the chosen DNS benchmark by

viii

Spalart & Strelets (2000). The nite-dierence solver is found to be dissipative


due to the use of a stability-enhancing lter. Its numerical dissipation is quantied and found to be comparable to the average eddy viscosity of the dynamic
Smagorinsky model, making it dicult to separate the eects of ltering versus
those of explicit subgrid-scale modeling. The negligible numerical dissipation of the
pseudo-spectral solvers allows an unambiguous assessment of the performance of
subgrid-scale models. Three explicit subgrid-scale models dynamic Smagorinsky,
, and truncated Navier-Stokes (TNS) are compared to a no-model simulation
(under-resolved DNS) and evaluated against the benchmark DNS data focusing on
two quantities of critical importance to airfoil and blade designers: time-averaged
pressure (Cp ) and skin friction (Cf ) predictions used in lift and drag calculations.
Results obtained with these explicit subgrid-scale models conrm that accurate
LES of laminar separation bubble ows are attainable with as low as 1% of DNS
resolution, and the poor performance of the no-model simulation underscores the
necessity of subgrid-scale modeling in coarse LES with low numerical dissipation.

ix

Chapter 1
Introduction
Laminar ow separation, transition to turbulence, and reattachment coined
laminar separation bubble because of the recirculation region the phenomena create
together directly aect the performance of increasingly important applications
ranging from small unmanned aerial vehicles to low pressure turbines found in
jet engines and gas generators. The presence of a laminar separation bubble on
a small unmanned aerial vehicles wing can increase the lift over drag ratio L/D
and thus the eciency of the craft. Airfoil shapes that maximize this eect are
sought at the design stage. On the other hand, laminar separation bubbles cause
detrimental unsteadiness and transient structural loads on low pressure turbine
blades, so blade shapes and ow control schemes to mitigate this phenomenon
are sought at the design stage. Faster, more accurate computational tools are
necessary to enable the design optimization of technologies to better control the
onset of laminar separation bubbles on blades and airfoils.

1.1

Laminar Separation Bubble Flows

The Reynolds number (Re) is a dimensionless ratio of inertial to viscous forces


characterizing the ow regime. Reynolds numbers for small unmanned aerial vehicles and micro-air vehicles are low to moderate. Based on wing chord length, they
are typically less than 2 106 and are in some cases only on the order of 104 to 105 .
By comparison, civilian airplanes are characterized by Reynolds numbers ranging

from a few million to 80 106 for the Boeing 747 at cruising velocity. Recent
experimental investigations of low Reynolds number aerodynamics by Hu et al.
(2007); Hain et al. (2009); Spedding & McArthur (2010) reveal that low to moderate Reynolds number ows over airfoils are often dominated by the eects of ow
separation and reattachment. The appearance of unsteady recirculation regions
due to separation and reattachment greatly inuence the aerodynamic forces the
wing is subjected to. They change the lift and drag characteristics and thus the
ight stability of small unmanned aircraft and micro aerial vehicles. Separationinduced transition is also seen in higher Reynolds number ows. For example,
ow separation and turbulent reattachment is sometimes observed on the blades
of wind turbines as well as on low pressure gas turbine blades due to the inuence
of the interior wall. This phenomenon causes unsteadiness in the ow, which is a
determining factor in high cycle fatigue of turbomachinery components. Although
most wind and gas turbines operating Re are higher than 105 , the physics governing separation-induced transition and the stability of the recirculation region it
creates are akin to those seen in laminar separation bubbles that occur on lower
Re applications like micro aerial vehicles.
A number of experiments have been performed to elucidate laminar separation bubble ows.

The seminal works by Gaster (1963) and Horton

(1968) established the foundation for the physical understanding of the laminar separation bubble and advanced some semi-empirical predictions for the
location of separation and reattachment.

Since then, a number of experi-

ments have been conducted to study in more detail the structure, stability,
and dynamics of laminar separation bubble ows Alving & Fernholz (1996);
Hggmark (2000); Marxen et al. (2003); Burgmann et al. (2006); Yarusevych et al.
(2006); Burgmann et al. (2007); Burgmann & Schrder (2008); Hu et al. (2007);
2

Figure 1.1: Features of the ow eld in the vicinity of a transitional separation


bubble Horton (1968); Lin & L.Pauley (1996); Castiglioni et al. (2014).

Hain et al. (2009); Spedding & McArthur (2010). As a result, the physical origin
of laminar and transitional ow separation is now qualitatively well understood.
As gure 1.1 illustrates, the attached laminar boundary layer developing on a wing
or blade is subjected to an adverse pressure gradient due to the airfoils curvature,
which causes it to separate. Immediately behind the separation point there is an
eectively stagnant ow region, the so-called dead air region. A reverse ow vortex develops downstream of the dead air region, leading to an inectional mean
velocity prole in the boundary layer. This triggers the growth of convective and
secondary instabilities which quickly break down to turbulence. Kelvin-Helmholtz
rolls are indeed visible in a snapshot of the ow eld of a laminar separation bubble
ow over a NACA0012 airfoil shown in gure 1.2. As the separated shear layer
transitions to turbulence, its interaction with the reverse ow vortex causes it to
3

Figure 1.2: Instantaneous iso-surfaces of spanwise vorticity on a NACA0012 airfoil


at 5 degrees angle of attack and Rec = 5 104 taken from a STAR-CCM+ large
eddy simulation Castiglioni (2015).
reattach, thereby closing o the recirculation region. Aft of the laminar separation bubble, unlike in the two dimensional case, the ow is fully turbulent and
three-dimensional. Clear spanwise vortices are not shed, but the size and shape
of the bubble changes in time due to the complex balance between the eects of
the pressure gradient, convective instabilities, and viscous dissipation. This picture emerges from these experimental investigations as well as from direct numerical simulations (DNS) results Lin & L.Pauley (1996); Spalart & Strelets (2000);
Alam & Sandham (2000); Marxen & Rist (2010); Jones et al. (2008, 2010).

Many DNS were carried out to match and augment experimental data. These
DNS shed light on the mechanisms of energy transfer at work in laminar separation
bubble ows Skote et al. (1998); Spalart & Strelets (2000); Skote & Henningson
(2002), the process of transition Na & Moin (1998); Wu & Moin (2010);
Alam & Sandham (2000); Marxen & Rist (2004), the stability characteristics of
such ows Marxen et al. (2003); Marxen & Rist (2010); Jones et al. (2010), and
the eect of disturbances and forcing on its dynamics Herbst & Henningson (2006);
Marxen & Henningson (2011); Jones et al. (2008). Agreement with experiments
was found to be very favorable in most cases. Key to getting agreement with
experimental results was resolving the reverse ow region near the wall, and the
shear layer which transitions to turbulence above the separation bubble.
Although three-dimensional laminar separation bubble ows have been shown
to dier substantially from their two-dimensional analog, there is some disagreement on the degree of importance of spanwise structures, their dominant wavelength, and the role they play in laminar separation bubble ows. Recent simulation results show that airfoil sections may require one chord length in the spanwise
direction to ensure the full three-dimensional nature of the laminar separation bubble can be captured Eisenbach & Friedrich (2008). Experimental studies conrm
that the dynamics and turbulent statistics of laminar separation bubble ows are
sensitive to incoming levels of turbulence, boundary conditions, and even acoustic
vibrations. This impedes the direct comparison of dierent experiments, and the
search for universal empirical parameters and predictors for its impact on lift and
drag as well as the level of turbulence it generates in the ow. Researchers devising
technologies to control or mitigate the onset of separation bubbles face the same
issues. A predictive tool fast enough to reliably explore the sensitivity to these
parameters and ow control schemes is needed.
5

1.2

Motivation

In order to produce more ecient airfoil or blade designs, to create control


schemes to reduce separation eects, and to better predict high cycle fatigue,
numerical prediction tools for laminar separation bubble ows are needed. Airfoil
and blade designers are primarily interested in obtaining accurate time-averaged
quantities relating to the aerodynamic forces on the airfoil or blade, namely the
coecient of pressure to measure the lift and drag, and the skin friction, another
important component of drag. To enable optimization in a realistic industrial
setting, these quantities must be calculated in a matter of a few hours or less. Such
computationally aordable, accurate and reliable numerical predictions for laminar
separation bubble ows had not been obtained until this study. As a proof of
concept and to avoid numerical issues associated with body-tted mesh generation,
an equivalent problem for ow over a at plate is formulated by imposing boundary
conditions that lead to a pressure distribution and Reynolds number based on
bubble length similar to those observed on the airfoils of small unmanned aircraft
and low-pressure turbine blades. The objective of this research is thus to test the
accuracy of large eddy simulation at resolutions drastically reduced compared to a
benchmark DNS for a laminar separation bubble ow over a at plate. Corollary
objectives are to identify key factors in obtaining accurate predictions, and to
investigate the performance of dierent subgrid-scale models in this highly underresolved environment.

Chapter 2
Background
2.1

Governing Equations

The relevant governing equations are the incompressible Navier-Stokes equations because the uid density and temperature may be considered constant for
laminar separation bubble ows at moderate Reynolds numbers (Re = 104 to
Re = 2 105 ).
p

ui
(ui uj ) =
+
+
t
xj
xi
xj
ui
= 0,
xi

ui uj
+
xj
xi

+ Fi ,

i = 1, 2, 3

(2.1)
(2.2)

The velocity eld u = (u1 , u2, u3 ) expressed in reference to a Cartesian coordinate


system x = (x1 , x2 , x3 ) is a solution of the momentum and continuity equations,
(2.1) and (2.2). p = P/ is the static pressure and is the kinematic viscosity
which is assumed to be constant and uniform in space.
No accepted analytical closed form solution to the Navier-Stokes equations for
laminar separation bubble ows is known. A numerical solution to the NavierStokes equations is thus sought on a discrete set of grid points xijk or discrete cell
volumes. A number of approaches exist to discretize the equations such as the
nite dierence, nite volume, and nite element methods to name only a few.
Despite fundamental dierences in their approach, each discretization methods
conservation properties are directly dependent on the accuracy of the schemes
7

used to approximate each of the terms in the equations solved. The accuracy of
the solution depends both on the exact formulation of the numerical scheme, the
quality (level of non-orthogonality) of the grid, and the level of resolution on which
the discretized equations are solved.

2.2

Direct Numerical Simulation

Once discretized, the Navier-Stokes equations are integrated in time. Doing


so without further approximations is called direct numerical simulation (DNS).
DNS is the most accurate and reliable uid dynamics simulation tool available.
However, to obtain an accurate solution using DNS requires that the mesh or
cell size captures all relevant scales of motion in the problem. If the ow to be
simulated is turbulent, then the mesh or cell size must be within one order of
magnitude from the Kolmogorov length scale, the smallest length scale at which
turbulence exists. Dominated by viscosity, these small scales are largely responsible
for the dissipation of turbulent motion into heat. Capturing Kolmogorov scales
imply solving the Navier-Stokes equations on a very large number of cells or a
very ne mesh. The computational work required may take a prohibitively large
amount of time even on the best available supercomputer. Indeed, DNS requires
substantial computational resources, long wall-clock runs, and long analysis times;
e.g. a relatively simple 3-D airfoil conguration at a Reynolds number of 5 104
required over 170 million grid points Jones et al. (2008). For laminar separation
bubble ow over a at plate at a Reynolds number of 105 , DNS results required
over 16000 processor-hours Cadieux et al. (2012). A number of 3-D congurations
and angles of attack need to be quickly investigated to allow for the optimization

of airfoil and turbine blade designs. For this case, a DNS approach is generally
impractical and other simulation approaches must be considered.

2.3

Reynolds-averaged Navier-Stokes

A widely-used computational uid dynamics simulation approach for moderate


to high Reynolds number ows is to solve for the mean ow quantities directly
instead of calculating the primary quantities at each instant in time as is done in
DNS. The primary quantities are split into uctuating and mean ow components
ui = Ui + ui

(2.3)

where Ui = hui i is the ensemble average of ui satisfying Reynolds averaging conditions. Substituting eq.(2.3) into the Navier-Stokes equations and noting that
huii = 0 gives the Reynolds-averaged Navier-Stokes (RANS) equations for incompressible ow:
Ui

1 P

+
(Ui Uj ) =
+
t
xj
xi xj
Ui
=0
xi
where

ijrans = hui uj i.

Uj Ui

+
xi
xj

ijrans

(2.4)
(2.5)
(2.6)

All information about the instantaneous velocity elds is lost through this averaging in favor of obtaining an estimate of the mean ow. Unfortunately, averaging
procedures used in practice, namely time and spatial averaging, do not strictly
satisfy Reynolds last condition that hhf (x)ig(x)i = hf (x)ihg(x)i for any ow that
is not fully developed with a large separation of scales, limiting its applicability to

such ows Wilcox (2006). In other words, if the mean and uctuating components
are correlated, then the time or spatial average of their product does not vanish
and the RANS equations are no longer valid Wilcox (2006). As seen in eq. (2.4),
the RANS equations have an additional unknown term, the Reynolds stress tensor
ijrans , which must be modeled using prior knowledge about the ow being simulated and/or additional equations. The majority of RANS models for ijrans rely on
the Boussinesq eddy viscosity approximation. The Boussinesq hypothesis states
that by analogy to momentum transfer in the molecular motion of a gas which
can be described by a molecular viscosity , the Reynolds stress tensor should be
proportional to the mean strain rate tensor using a turbulent eddy viscosity rans :

ijrans rans

Uj
Ui
+
xi
xj

(2.7)

This reduces the number of extra unknowns from six to one, the eddy viscosity.
Most RANS models use dimensional arguments and analogy to other physical
processes to set the value for rans . For example, Prandtls mixing length model
uses the Boussinesq analogy to molecular momentum transport and assumes that
there exists a turbulent mixing length analogous to the mean free path Wilcox
(2006). Using dimensional analysis, the mixing length model for a boundary layer
simplies to
U
,
y
U
|,
= 2mix |
y

rans = rans

(2.8)

rans

(2.9)

mix = =

q
U(y)
dy 1.72x/ Rex ,
1
U

(2.10)

10

where the mixing length is assumed to be proportional to the displacement thickness of the boundary layer. Such a simple turbulence model provides results in
agreement with experiments for boundary layers Wilcox (2006). However, it is
incomplete because the appropriate mixing length for the ow being simulated
must be known a priori. Moreover, the mixing length model is only valid for simple ows with slow-varying properties (so-called equilibrium turbulent ows) due
to the assumptions used in deriving it. Modern models are far more complex and
perform better in a wider range of ows, but still struggle with accurately predicting transition from laminar to turbulent ow and the lengths of recirculation
zones.

2.3.1

RANS for Laminar Separation Bubble Flows

Laminar separation bubble ows present a challenge to RANS models because


boundary layer separation and reattachment involves subtle interactions between
viscous, advective, and pressure eects and is inherently a non-equilibrium process,
especially when driven by an adverse pressure gradient instead of geometry (e.g.
backward facing step). Although the rate of strain changes rapidly as the ow
separates, the turbulence adjusts to changes in the ow on an unrelated, longer
time scale. A perturbed turbulent boundary layer was experimentally shown not
to return to equilibrium for at least 10 boundary-layer thicknesses downstream of
the perturbation Wilcox (2006). Since the Reynolds stresses modeled in typical
two-equation RANS models are adjusted based solely on the mean rate of strain,
they preclude any such transient eects from ow history. Attempts were made to
adjust for this eect by relaxing the eddy viscosity in the region behind separation.
However, this method requires prior knowledge of the separation point - and thus

11

precludes strong predictive capabilities even for unsteady RANS Howard et al.
(2000).
Spalart & Strelets (2000) tested a number of typical RANS turbulence model,
namely Spalart-Allmaras, Menters shear stress transport, modied shear stress
transport, and Secundovs t92 , for a simple laminar separation bubble ow on
a at plate driven by suction from the top. All RANS turbulence models except
Spalart-Allmaras predict earlier transition and reattachment than observed in their
spectral DNS results. Obvious disagreement in the location and magnitude of the
peak negative skin friction is observed between the results of dierent models.
All RANS results were found to under-predict the level of skin friction downstream of reattachment. Without modication, Menters shear stress transport
model transitions even before the expected separation point, predicting attached
ow throughout the domain. Although Spalart-Allmaras results were encouraging,
none of the RANS turbulence models Spalart & Strelets (2000) tested recovered
all the important features of the time-averaged skin friction DNS curve: location
of the separation point, location and magnitude of the peak negative skin friction,
reattachment point, and turbulent skin friction levels immediately downstream of
reattachment. The large variations observed in skin friction predictions depending
on the RANS turbulence model chosen may be symptomatic of a lack of robustness of the typical RANS turbulence modeling approach when applied to inherently
unsteady phenomena.
To better predict transition, Howard et al. (2000) sensitized two-equation
RANS model coecients of the Launder and Sharma and k g models to the local
turbulent Reynolds number as proposed by Wilcox (2006), removing the need for
a priori knowledge of the separation point. With the local turbulent Reynolds

12

number sensitivity modication, two-equation models were shown to predict transition more reliably in an unsteady RANS solver and showed improved agreement
for separation bubble length. However, Launder and Sharma and k g model
results were still inconsistent amongst each other and amongst dierent sensitization approaches. None matched the DNS results for peak negative skin friction
and its level immediately downstream of reattachment Howard et al. (2000).
RANS turbulence models using second moment closures show improvement
over single and two-equation models without the need for articially triggering
transition, but require tuning of their closure coecients for lower Reynolds number ows with separation and transition Hadi & Hanjali (2000). While results
show better agreement for the reattachment point and turbulent region at coarser
resolution than the Spalart-Allmaras model, second-moment closure RANS results
still fail to recover the peak negative skin friction. Since the authors mention but
do not present results for a ner grid, it can be surmised that results did not
improve signicantly Hadi & Hanjali (2000). That the method does not converge to DNS results with increased resolution is further evidence that it is not
well-suited for the problem.
Similarly, results for a laminar separation bubble on a at plate with a semicircular leading edge using a two-layer model matched experimental results reasonably well, but required a number of empirical correlation along with special modications to the model to capture transition Papanicolaou & Rodi (1999). Although
RANS can predict separation and transition reliably with models optimized for
such ows (second-moment closure, two-layer model), it still struggles to recover
wall skin friction accurately for this proof-of-concept laminar separation bubble
ow over a at plate.

13

2.4

Large Eddy Simulation

Another option is to employ large eddy simulation (LES) techniques. LES techniques were developed based on the observations made by Kolmogorov that the
smallest scales of turbulence are dominated by viscosity, behave mostly isotropically and account for the majority of turbulent dissipation. The understanding of
the energy cascade from large to small scales led to the idea of modeling the more
universal small scales of turbulent motion while resolving the energy containing
eddies directly aected by the ow boundary conditions. This greatly relaxes the
DNS requirement that the mesh or cell size be on the same order as the Kolmogorov
length scale. LES hinges on the use of subgrid-scale models to predict the correct
small scale dissipation rate based solely on information from the larger resolved
scales.

2.4.1

Filtered Navier-Stokes Equations

A low-pass lter operation is used to separate small scales from larger ones
and derive the ltered Navier-Stokes equations. This ltering operation can be
described by a convolution integral with the lter function or kernel G,
ui (x, t) = G ui =

G(x x ; )ui (x , t) d3 x

(2.11)

where the resolved scale or ltered velocity is ui , and the subgrid-scale velocity
is dened as ui = ui ui . is the lter width which is generally taken to

14

be proportional to the grid or cell size = (xyz) 3 Sagaut (2006). The


incompressible ltered Navier-Stokes equations can then be written as follows:
p

ui
(
ui uj + ijsgs ) =
+
+
t
xj
xi
xj
ui
= 0,
xi

ui uj
+
xj
xi

+ Fi ,

(2.12)
(2.13)

where
ijsgs = uiuj ui uj

(2.14)

is the subgrid-scale stress tensor. It contains the term ui uj which is a new unknown.
Just as turbulence models are required to close the RANS equations, the ltered Navier-Stokes equations require a subgrid-scale model for ijsgs . However,
the derivation of subgrid-scale models generally rely only on the assumption that
the eect of the scales of motion smaller than the lter width on the larger scales
are small and mostly dissipative. This assumption is more robust than those made
in deriving the RANS equations and most RANS turbulence models because it is
satised for a much wider array of ow conditions and relative lter widths.

2.4.2

Subgrid-scale Models

There exists a variety of dierent subgrid-scale models to close equations (2.12)


and (2.13) that generally belong to one of two distinct categories: structural and
functional modeling. Structural modeling directly approximates the subgrid-scale
stress tensor or subgrid-scale velocities based on the resolved velocities or a formal
series expansion. This approach assumes that the structure of the small scales
is universal and the energy contained in the subgrid-scales are a function of the
15

resolved scales Sagaut (2006). Examples include models based on the deconvolution procedure, stress transport models, and subgrid-scale velocity reconstruction
models. A recent example is the velocity estimation model by Dubois et al. (2002).
Instead of directly approximating ijsgs , functional modeling seeks to approximate the eects of inter-scale energy transfer on the resolved scales. Rather than
assuming that the structure of the small scales is universal, this approach posits
that the eects of the small scales on the larger resolved scales are universal Sagaut
(2006). Knowledge of the turbulent energy cascade and the concepts of forward
and backscatter are used to justify the assumption that the eect of the small
scales on the large is universal and depends only on the energy of the large scales
driving the ow. Recent examples of such subgrid-scale models are the and the
interscale energy transfer models Nicoud et al. (2011); Anderson & Domaradzki
(2012). The increasingly popular implicit LES (ILES) approach also generally falls
into this category, where the numerical scheme is adjusted such that its truncation
errors and associated dissipative and dispersive eects have the desired impact on
the resolved scales. The most famous functional modeling example remains the
Smagorinsky model.
The Smagorinsky Model
The Boussinesq approximation used in many RANS models is invoked again,
but this time to calculate a turbulent eddy viscosity sgs describing the subgridscale dissipation based solely on the resolved scales of motion.
1 sgs
1
ijsgs kk
ij = 2sgs (Sij Skk ij ),
3
3

(2.15)

16

where
Sij ,
sgs = (CS )2 |S|
= (2Sij Sij ) 2
|S|

(2.16)

1
Sij =
2

(2.17)

ui uj
+
xj
xi

= (xyz) 3 .

(2.18)
(2.19)

CS is a closure coecients determined ahead of simulation by matching available experimental or DNS data. This is a limitation to the models applicability
because foreknowledge of the ow being simulated is required to set the closure
coecient. It also cannot account for the local reduction of eddy viscosity near
physical boundaries without the use of explicit Van Driest damping functions. A
modication of the Smagorinsky model which eliminates these issues and makes it
universal consists of letting the closure coecients be functions of time and space
(e.g. CS2 = Cd (x, t)). In this dynamic version of the Smagorinsky model, these
coecients are computed dynamically using the Germano identity.
The Dynamic Procedure
The dynamic procedure used to compute the local instantaneous closure coecient Cd (x, t) for the Smagorinsky model is used here as an example, but it can be
applied to other models formulated in a similar fashion. Its purpose is to provide
this coecient based on local resolved strain rate without prior knowledge of the
ow. It eectively reduces the model contribution in laminar shear ows where

17

the stress tensor is not zero, but no turbulence exists, removing the issues that the
static Smagorinsky model faces with physical boundaries.
Sij
sgs = Cd 2 |S|
Cd = min 0.22 ,

(2.20)
max

"

hLij Mij i
,
hMij Mij i

#!

b
b
Lij = ud
j u
i u
j ,
iu

(2.21)
(2.22)

b b
b 2 |S|
2 |S|
S 2
Mij = 2\
S ij ,
ij
1

b = 2 = 2(xyz) 3 .

(2.23)
(2.24)

The overbar is used to represent grid-ltered terms (lter width ) and the hat is
used to indicate test-ltered quantities using Simpsons rule (lter width 2):
2
1
1
fb(x) f (x x) + f (x) + f (x + x).
6
3
6

(2.25)

The weights of the lter are adapted to the non-uniform vertical grid using
quadratic interpolation. The h i symbol used in (2.21) denotes averaging in any
uniform direction (if one exists) or local averaging in all directions. Although small
negative values of Cd may be physically justied, mimicking backscatter phenomena, such negative values can quickly lead to numerical instability. So in practice,
averaging is used in tandem with clipping to avoid negative as well as rapidly
oscillating values of Cd . The constant computed through this dynamic procedure
is local in space and time. It eectively reduces the model contribution in laminar shear ows where the stress tensor is not zero, but no turbulence exists. The
dynamic procedure is particularly computationally expensive due to the application of a test-lter in three-dimensions on at least two tensorial quantities ud
j and
iu

2 |S|
S as well as local spatial averaging, each ltering and averaging operation
\
ij

18

often requiring communication among dierent processes or blocks in parallelized


implementations. Despite this shortcoming, the dynamic Smagorinsky model has
become the benchmark against which other subgrid-scale model are tested due to
its success in academia and its universality.
The Model
The model follows denitions set forth in (2.15) and (2.18) but computes
sgs using the singular values i of the velocity derivative tensor gij . This choice is
motivated by the desire to improve on the dynamic Smagorinsky model by providing more appropriate near-wall behavior, as well as providing zero contributions
in pure two-dimensional shear or pure rotation cases Nicoud et al. (2011).
sgs = (C )2

3 (1 2 )(2 3 )
12

is the subgrid-scale eddy viscosity,


(2.26)

where
1 2 3 0,
C = 1.35

are the singular values of gij =


is the closure coecient.

ui
, and
xj

(2.27)
(2.28)

C is determined from homogeneous turbulence and validated using channel ow


simulations Nicoud et al. (2011). The singular values i are obtained using the
invariants of Gij = gkigkj and their angles to avoid the overhead of linear algebra
library calls to an eigenvalue solver for each cell in the domain at each time step.

19

The truncated Navier-Stokes Approach


The truncated Navier-Stokes approach follows the method developed by
Domaradzki et al. (2002) in which periodic ltering is used as a substitute for
a subgrid-scale model. Periodic ltering is used to remove energy from the smallest resolved scales by the use of a low-pass approximate deconvolution method
(ADM) lter Stolz et al. (2001). The ltering operation is implemented using
the product of an approximate deconvolution lter QN G1 described in
Tantikul & Domaradzki (2010) with lter G:
QN G = I (I G)N +1 .

(2.29)

The order N = 5 is chosen such that the lter only aects scales smaller than lter
width = x when using a simple three point lter in physical space G(x):
1
3
1
G(x) f (x) f (x x) + f (x) + f (x + x).
8
4
8

(2.30)

The lter weights are adjusted for the non-uniform vertical grid using quadratic
interpolation. Since the energy cascades from large to small scales and accumulates
there slowly in a high order under-resolved simulation, it is only necessary to lter
after a fraction of a percent of the large eddy turnover time. Hence, ltering is only
applied when the kinetic energy at high wave numbers reaches unphysical levels
in the truncated Navier-Stokes approach. This is fundamentally dierent from
implicit LES, where the inherent approximation errors can be said to act as a lowpass lter at each time step. Excessive energy accumulating in the small scales is
detected using a criterion based on the ratio of energy removed I(x)/I(2x) by

20

two ADM test-lters with lter widths = x denoted by the tilde and = 2x
denoted by the hat as follows
E E
dV
b
V E E
+
Z Y * P3 1
i)(ui ui)
i=1 2 (ui u

(y)dy
P3 1
b i )(ui u
bi)
0
i=1 2 (ui u

I(x)
=
I(2x)

(2.31)
(2.32)

ui = (Q5 G(x)) ui

(2.33)

ubi = (Q5 G(2x)) ui

(2.34)

1
1
1
G(2x) f f (x x) + f (x) + f (x + x).
4
2
4

(2.35)

The ratio I(x)/I(2x) represents the energy contained in the small scales compared to the larger scales. When it reaches values in excess of those obtained
for a typical dissipation, inertial, or Batchelor energy spectrum determined to
be 0.007 to 0.009 from theory by Tantikul & Domaradzki (2010, 2011) primary
variables are ltered in physical space with lter Q5 G(x). Using this criterion,
the lter is applied at varying intervals centered around 200 hundred time steps for
the coarsest resolution simulation presented here, corresponding to approximately
0.5% of one non-dimensional time unit t = t LUx0 .

2.4.3

LES Results for Laminar Separation Bubble Flows

A number of LES of laminar separation bubble ows over at plates and


airfoils have been completed recently by Wilson & Pauley (1998); Yang & Voke
(2001); Roberts & Yaras (2005); Eisenbach & Friedrich (2008); Xu et al. (2010);
Kojima et al. (2013). For instance, LES results Yang & Voke (2001) obtained with
the dynamic Smagorinsky model were reported to be in good agreement with experiments for boundary-layer separation and transition caused by surface curvature
21

at Re = 3, 450. Yet even for this relatively low Reynolds number, the two critical
issues in getting agreement were a numerical resolution (4727264 mesh points)
comparable to DNS of the same ow, and a high order numerical method. Such
strict requirements are dicult to satisfy in simulations of practical ows often
performed with low order nite dierence or nite volume methods (e.g. commercial codes). Similarly, LES of ow separation on an airfoil at a high angle of attack
was performed at Re = 105 using Cartesian grids Eisenbach & Friedrich (2008).
This case also required very high resolutions between 50 and 100 million mesh
points to obtain good agreement. Using LES with such high resolution and higher
order methods implies a time-to-solution on the same order as DNS. Therefore,
the question remains: can LES produce suciently accurate results for laminar
separation bubble ows with drastically reduced resolution, around 1% of DNS
resolution, commonly achievable for fully turbulent ows?
While a handful of other investigators also performed low resolution LES, the
eect of dierent subgrid-scale models on the quality of important results such as
time-averaged skin friction and pressure coecient remains largely unknown for
laminar separation bubble ows. Only the constant Smagorinsky model has been
compared to its dynamic counterpart and a no-model case in Wilson & Pauley
(1998).

Other investigators like Eisenbach & Friedrich (2008); Yang & Voke

(2001); Xu et al. (2010) relied entirely on the dynamic Smagorinsky model, or


on the numerical dissipation of their chosen scheme as in implicit LES or ILES
in Kojima et al. (2013) and even without any prior knowledge of the dissipative
schemes eects on the resolved scales in the case of Roberts & Yaras (2005). Without benchmark DNS data or a baseline case with no subgrid-scale model active to
compare directly to, the performance of their models or ILES results could not be
evaluated quantitatively.
22

2.5

Wall-modeled LES and Hybrid RANS-LES

For LES to be accurate, it requires a mesh nearly as ne as a DNS near physical boundary where boundary layers develop. To mitigate this stringent resolution
requirement, wall models are developed to give approximate boundary conditions
to the LES solver away from the surface. A wide variety of models with dierent assumptions have been proposed and reviewed by Piomelli & Balaras (2002);
Piomelli (2008) and Sagaut & Deck (2009). The use of wall models generally
mitigates noise generated by poor approximation of curved surfaces or highly nonorthogonal body-tted meshes in low resolution settings and permit simulations to
reach much higher Reynolds numbers that are closer to operating conditions for
most turbomachinery blades McMullan & Page (2012). However, wall models have
historically had a poor track record in predicting separation and reattachment, and
introduce another source of error into LES due to further approximations made at
the wall Bose & Moin (2014). But more importantly, when wall models are used
it becomes dicult to distinguish between the performance of the wall model and
that of the subgrid-scale model because obtaining correct amount of turbulent content near the wall is key to the overall accuracy of LES. Detached eddy simulation
(DES) solves the RANS equations near the wall and smoothly transition to LES
on a single grid using a single hybrid RANS-LES turbulence model developed by
Spalart (2006). This removes the need for wall-layer modeling while still drastically
reducing near-surface resolution requirements, allowing the simulation of higher
Reynols number ow. Despite many successes, dicult issues such as modeled
stress depletion in the log law region, and non-monotonic grid convergence arise
in DES as well as its derivatives delayed DES and zonal DES as acknowledged by
Spalart (2006, 2009); Deck et al. (2011); Deck (2012). Partially averaged NavierStokes (PANS) developed by Girimaji & Abdol-Hamid (2005); Basara et al. (2011)
23

and other variable resolution approaches such as the scale-adaptive simulation


(SAS) proposed by Menter & Egorov (2010); Egorov et al. (2010) and turbulenceresolving RANS (TRANS) put forth by Shur et al. (2008) all avoid these issues by
tying the eddy viscosity to physical quantities like energy grid density or an integral
length scale. However, their turbulence modeling approaches and by consequence
their results are often closer to unsteady RANS than LES and are thus limited in
their ability to capture transient and unsteady eects accurately as pointed out in
Menter & Egorov (2010). In fact, neither hybrid approaches nor wall models have
been validated to the same extent as LES, even when in pure RANS or pure LES
mode (when the method allows it) Sagaut & Deck (2009). To investigate the eects
of dierent subgrid-scale models without the unknown inuence of wall models or
hybrid RANS-LES approaches, wall-resolved LES is chosen for this work.

24

Chapter 3
Methodology
3.1

Approach

Laminar separation bubble ows occur on blades and airfoils at low to moderate Reynolds numbers ranging from 104 to 105 due the curvature of the airfoils
and blades. Simulating ow over blades and airfoils requires the creation of nonorthogonal body tted meshes, unstructured grids, or the use of immersed boundary methods to properly represent the airfoil or blades surface. Grid creation not
only presents its own challenges, but also often limits numerical solution methods
to second order accuracy in space and time, with the exception of nite element and
discontinuous Galerkin methods. The choice of meshing technique may also have
inherent approximations that in turn aect stability and accuracy of the numerical methods at low resolution. For example, approximating a curved surface by a
series of connected straight lines as opposed to bezier curves or splines can give rise
to inaccurate results in second order codes, and catastrophic numerical instability
when using higher order methods.
In order to investigate the capacity of LES to reduce the resolution requirement for laminar separation bubble simulations and the performance of dierent
subgrid-scale models free from the numerical issues associated with geometry, ow
over a at plate with an adverse pressure gradient strong enough to cause separation as described in section 3.2 is studied. This approach has been used sucessfully
to study laminar separation bubble ows both in experiments by Hggmark (2000);
25

Marxen et al. (2003); Sohn et al. (1998) and in simulations by Spalart & Strelets
(2000); Alam & Sandham (2000); Wilson & Pauley (1998); Herbst & Henningson
(2006); Na & Moin (1998); Skote et al. (1998); Wu & Moin (2010). The resulting
pressure distribution is qualitatively comparable to what is seen on blades and
airfoils as is shown in gure 3.1: a smooth increase in pressure is followed by a
plateau over the separation bubble. The plateau ends with a sharp rise in pressure
indicating the ows transition to turbulence and reattachment. Downstream of
the sharp rise, the pressure plateaus again over the developing attached turbulent
boundary layer. The only dierence to note between the airfoil and the at plate
laminar separation bubble pressure distribution is that the sharp peak in pressure
at the stagnation point of the airfoil is not observed on the at plate. Despite
this dierence, the Reynolds number based on bubble length (Re 67, 000) is
similar to those found on airfoils and blades indicating a degree of physical equivalence. In fact, at plate laminar separation bubble ows display the same physical
features as those seen on airfoils as evidenced by the similarity of the mean velocity contours and mean velocity proles in gure 3.2 to those in gure 1.1. The
mechanisms for separation-induced transition are also the same as in the airfoil
case. Notice that the Kelvin-Helmholtz rolls visible on a at plate with suction
from the top shown in gure 3.3 near x = 3.5 closely resemble those seen on an
airfoil shown in gure 1.2, each gure displaying iso-surfaces of spanwise vorticity.
As such, general conclusions reached from investigating at plate separation and
reattachment should also be applicable to blades and airfoils.

26

3.2

Flow Specification

The computational setup used by Spalart & Strelets (2000) to study separationinduced transition ow over a at plate is followed. The physical domain is a
rectangular box with height Y , length 7.5Y , and width 0.6Y (see gure 3.4). At
the inow a laminar Blasius boundary layer velocity prole is imposed with the
free stream velocity U0 . At the top boundary, a vertical suction velocity is imposed
in a narrow slot oriented perpendicular to the mean ow direction. The suction
produces an adverse pressure gradient that causes ow separation. The ow then
transitions to turbulence and reattaches downstream. The vertical suction velocity
is specied as
V (x) = a exp([(x xs )/(0.24Y )]2 ),

(3.1)

where a is the peak velocity and xs is its streamwise location Spalart & Strelets
(2000). The resulting separation bubble is sensitive only to the upper-wall boundary conditions through the nominal ow deceleration parameter S,
1
S=
Y U0

V (x)dx.

(3.2)

Using the height Y to non-dimensionalize all relevant lengths the parameters in


the equations above are set such that xs = 3, S = 0.3 and the Reynolds number
at xs is Rexs = 105 , giving a 0.7U0 and ReY = Rexs /3, matching those in
Spalart & Strelets (2000). These choices are driven by the requirement that the
ow separates naturally, without additional forcing mechanisms like those used in
Alam & Sandham (2000).

27

3.3

Numerical Methods

The numerical schemes used to approximate the equations in space and integrate them in time have an impact on three important quantities: the rate of convergence, numerical dissipation and dispersion. The choice of numerical scheme
determines the rate of convergence to the true solution. For example, a second
order scheme in space implies that doubling the number of mesh points over a
given solution domain should decrease the error of the simulation by a factor of 4
over a constant time of integration.
The concepts of numerical dissipation and dispersion are linked to the exact
form of the truncation error terms of any approximation made by the scheme. For
example, a spatial derivative approximated by a central dierence has a truncation
error E calculated from its Taylor series expansion as follows.
f (x + h) f (x h)
+ h2 x3 f (x) + O(h3)
2h
f (x + h) f (x h)

2h

x f (x) =

E = h2 x3 f (x) + O(h3)

(3.3)
(3.4)
(3.5)

The omission of these higher order terms in the simulation have eects that are
unknown a priori and depend on the governing equations, the particular ow simulated and the degree of under-resolution. A scheme is described as dissipative if its
total energy kinetic decreases faster than the exact solution. Its visible eect can be
likened to articially increasing viscosity. The classic example is the reduction in
amplitude of a half sine wave, and its increasing wave length during linear convection in space using a simple nite-dierence upwind scheme as seen in gure 3.5a
reproduced from Fletcher (1991). Numerical dispersion is linked to the scheme

28

amplifying and attenuating dierent Fourier modes of a derivative approximation


causing oscillations that travel with dierent wave speeds. For example, the same
half-sine wave linear convection problem solved using a Crank-Nicolson scheme
results in spurious oscillations as seen in gure 3.5b reproduced from Fletcher
(1991). Such spurious oscillations are characteristic of a scheme with a dominating dispersive term. Dissipative schemes have been preferred historically for their
robustness: their ability to articially smooth out any sharp changes or discontinuities in solutions where less dissipative or more dispersive schemes might become
numerically unstable and not provide any results. This is problematic for three
reasons. Solutions for ows with shocks where discontinuities are physical will be
increasingly inaccurate over time. For ows where transition to turbulence occurs
naturally, as in laminar separation bubble ows, excessive dissipation may inhibit
transition to turbulence and reattachement completely. Finally, any highly underresolved simulation with a dissipative scheme will likely be particularly inaccurate.
Under-resolution already implies that not all length scales of motion relevant to
the problem will be captured. The amount of numerical dissipation is generally
inversely proportional to the resolution the more under-resolved, the higher the
numerical dissipation. Combined, under-resolution and dissipative schemes may
even preclude the development of high wave number content in primary quantities.
Since most LES models are predicated on the ability to predict the correct subgridscale dissipation rate based solely on coarser resolved scales, energy conservation
in the numerical methods used is paramount as evidenced in Kravchenko & Moin
(1997). Second-order methods often damp and deform high wave number content of the primary variables. Subgrid-scale models are generally not designed to
account for these eects, and given such awed input are unlikely to compute the

29

correct subgrid-scale dissipation. As such, understanding, controlling, or removing numerical dissipation at low resolution is of critical importance in predicting
laminar separation bubble ows accurately and quickly. For these reasons, two
Navier-Stokes solvers that employ high-order numerical methods were chosen to
investigate the capability of LES to reduce the resolution requirements for accurate
laminar separation bubble predictions.

3.3.1

Center for Turbulence Research Code

Developed by graduate students at the NASA Center for Turbulence Research


(CTR) at Stanford, this code solves the compressible LES equations for a perfect
gas Nagarajan et al. (2007). Henceforth this solver will be referred to as the CTR
code. Derivatives are computed using a sixth-order nite dierence approximation
similar to a Pad scheme. The free parameters are chosen such that the resulting
derivative approximations resolve higher waves numbers than otherwise possible
Lele (1992). These high wave numbers are generally not well approximated or
even severely damped in standard nite dierence schemes of the same order.
To maintain the spectral-like eciency and high order of convergence of these
derivative approximations, the horizontal directions are treated as periodic and
sponge regions are used to simulate non-periodic ow. A split implicit-explicit
time integration is used. For explicit time advancement in the freestream ow,
a third-order Runge-Kutta scheme (RK3) is employed. A second-order A-stable
implicit scheme is used near the wall to allow for larger stable integration time
steps. Compact ltering as described in Lele (1992) is employed at each time step,
both in the freestream and wall-normal directions. Filtering is necessary to remove
aliasing errors, for overall stability and to ensure a smooth continuous solution at
the interface of the implicit and explicit computational domains Nagarajan (2004).
30

The numerical scheme is constructed on a structured curvilinear grid, and the


variables are staggered in space. The freestream Mach number is chosen to be 0.2.
For compressible ows subgrid-scale models are developed in terms of Favre-ltered
quantities, as addressed for the rst time in detail by Erlebacher et al. (1992),
and in this work the dynamic Smagorinsky model is used in a form described in
Sayadi & Moin (2012).
Due to the periodic boundary conditions in the streamwise direction, numerical
sponges are necessary to simulate spatially evolving ow. This sponge region allows
the ow to be recycled from outlet to inlet by forcing a return to the desired inlet
boundary layer prole. Due to the code being compressible, a numerical sponge
at the top boundary is also necessary to ensure sound and vortical waves are not
reected back into the computational domain Mani (2012). The inlet sponge region
spans from x = 0.03 to x = 0.5, whereas the outlet sponge starts at x = 8 and ends
at x = 9.2. The top sponge extends the domain from y = 1 to y = 1.8. Sponge
regions account for one third of the total number of mesh points. These sponge
layers relax the computed Navier-Stokes solution to the scale-similar compressible
boundary layer case obtained a priori as a reference solution. From y = 1 to
y = 1.4, the reference solutions wall-normal velocity is changed to the suction
prole specied in eq. (3.1). It is then smoothly brought back to its precomputed
scale-similar value from y = 1.4 to y = 1.8. The sponge relaxation parameter
increases from zero at y = 1, the end of the physical domain, and reaches its
maximum close to the end of the sponge region at y = 1.8. The suction velocity is
thus enforced indirectly through the inuence of the forced solution above the top
of the physical domain.

31

3.3.2

Spectral code in vorticity form

A pseudo-spectral incompressible Navier-Stokes solver originally developed by


Domaradzki & Metcalfe (1987) was modied and parallelized to perform LES of
laminar separation bubble ow. The derivatives in the horizontal directions are
approximated using Fourier expansions, whereas the vertical is approximated using
collocated Chebyshev polynomials. The ow variables are integrated in time using
a fractional time step method described in Orszag & Kells (1980). It solves the
ltered Navier-Stokes equations in rotational form:
ui

+
= [ijk uj
k ]
t
xi
xj
ui
= 0,
xi

ui uj
+
xj
xi

+ Fi

ijsgs
,
xj

i = 1, 2, 3

(3.6)
(3.7)

where

i = ijk

uk
xj

is the vorticity,

(3.8)

= P/ + 1/2 ui ui

is the pressure head,

(3.9)

Fi = fi (x, t),

is a body force, and

(3.10)

is the subgrid-scale stress tensor.

(3.11)

ijsgs = ui uj uiuj

Strong algebraic grid stretching is used in the vertical to redistribute the Chebyshev
grid points such that approximately 2/3 of the points lie inside the boundary layer at
its thickest while still resolving the inow Blasius boundary layer. The streamwise
velocity is then decomposed into a base inow prole and velocity defect as follows:
+ u, where u
= [
utotal = u
u(y), 0, 0] is known and computed to be the Blasius

32

solution at the inlet x = x0 . Hence, u is the departure away from the inow prole
at each location in space that is solved for at each time step.
The non-linear term in u is computed in rotational form as in eq. (3.6). It is
advanced in time using Adams-Bashforth scheme, whereas the advection due to the
is computed separately using Crank-Nicolson to obtain better stability
base ow u
and kinetic-energy-preserving characteristics as seen in eq. (3.12) and (3.13).
"

ijsgs
3

n
ui = ui + t Ni + Fi
2
xj
!

ui
ui
1

+
ui = ui + uit
2
xj
xj

#n

"

ijsgs
1
t Ni + Fi
2
xj

#n1

(3.12)
(3.13)

where

(3.14)

Ni = ijk uj
k ui

ui
xj

represent the non-linear term contributions in ui.


(3.15)

The base ow advection, pressure and viscous steps are performed using the Fourier
representation of the ow variables
ui(x, y, z, t) =

ui(kx , y, kz , t) exp(ikx x) exp(ikz z),

(3.16)

|m|<M |n|<N

where kx = 2m/Lx and kz = 2n/Lz are the horizontal wavenumbers in the


corresponding streamwise x and spanwise directions z, and Lx and Lz are the
periodicity lengths. The base ow advection step described in eq. (3.13) then
simplies to eq. (3.17).
(1 + b(y)kx )
u ,
u
i =
(1 b(y)kx ) i

1
where b(y) = u(y)t
2

(3.17)

33

The pressure step consists of a correction to ensure the resulting ow eld is


divergence-free where the pressure is treated implicitly. Used together with the
continuity equation (3.7), this results in a Poisson equation for vertical velocity
with known Dirichlet boundary conditions as shown in eq. (3.18). The scheme
of Orszag & Kells (1980) removes the need to impose explicit pressure boundary
conditions at the expense of a small numerical error near the wall which is O(t).
If a consistent boundary condition for the vertical velocity is used as in eq. (3.22),
3

O(t 2 ) may be reached Guermond et al. (2006).




k 2 v ,
D 2 k 2 v = D[ikx u + ikz w]

(3.18)

where

(3.19)

,
y

(3.20)

k 2 = kx2 + kz2 ,

(3.21)

D=

The above Poisson equation (3.18) is solved with Dirichlet boundary conditions
h

v |b = v |b t D (ikx (u un ) + ikz (w w n )) + k 2 (v v n ) |b (3.22)


such that the pressure head and horizontal velocities can then be obtained algebraically as follows
= [D v + ikx u + ikz w ]/(k 2 t)

(3.23)

u = u ikx t

(3.24)

w = w ikz t

(3.25)

34

The viscous terms are then treated implicitly, which results in one Helmholtz
equation for each direction in spectral space as shown in eq. (3.26).


1
2
un+1
=
(u + ui )
D k
i
t
t i
2

(3.26)

Numerical sponge regions are implemented in the streamwise direction using the
fringe method formulation of Spalart & Watmu (1993) to damp all turbulence
and return the outow to the desired inow Blasius prole u(y) using the body
force term (3.27):
Fi = (x)(
ui ui),

(3.27)

where


(x) = f e((xx0 )/l ) + e((xxmax )/r )

f = 10
l =

U0
,
Lf

1
Lf ,
10

tends to zero outside inow and outow,


(3.28)
is the forcing strength,

3
r = Lf ,
8

(3.29)

are the widths of sponge regions,


(3.30)

Lf =

15
(xmax x0 )
100

is the total size of the sponge regions.


(3.31)

This forcing formulation enables the simulation of a spatially evolving boundary


layer while maintaining periodic boundary conditions and spectral accuracy. The
sponge region at the inow extends from x = 0.25 to x = 0.45, whereas the sponge
at the outow spans x = 8.7 to x = 10. Fringe method forcing terms are active

35

over 15% of the total domain. Their inuence does extend somewhat beyond where
the terms are active, so to be safe the region x = 0.5 to x = 7.5 is considered to
be the physically representative region of our computational domain.
A ceiling suction boundary condition that matches Spalarts and is designed
to cause a 30% free stream ow deceleration are enforced. The resulting top
streamwise velocity is computed during the pressure step by applying the continuity
equation and imposing zero spanwise velocity. In practice, the eective deceleration
is closer to 25% due to viscous eects as reported by Spalart & Strelets (2000).
To ensure mass is conserved in the complete domain, the amount of uid removed
by suction from the top of the domain is injected in a narrow slot outside of
the region of interest (from x = 8.5 to x = 9.5). The streamwise and vertical
velocity boundary conditions are shown in gure 3.6. Blowing is applied at the
wall to avoid any stability problems associated with disturbances near the top of
the domain where the resolution is very coarse. Blowing through the plate severely
limits the CFL number used in our simulations, increasing overall computational
time. The upshot of this strict CFL restriction is that it removes any possibility
of implicit ltering due to large time steps, a phenomenon observed in previous
results using implicit schemes.

3.3.3

Spectral code in skew-symmetric form

The vorticity form of the discretized Navier-Stokes equations was shown


to be prone to aliasing errors and more sensitive to Gibbs phenomena by
Kravchenko & Moin (1997), which can lead to persistent numerical oscillations at
the smallest resolved scales in a spectral code due to the discretization methods
inherently negligible numerical dissipation. Concern that these sources of error

36

might be exacerbated at low resolution and aect the ability of explicit subgridscale models to provide the correct eddy viscosity led to the development of a spectral code with a higher order time integration scheme where the non-linear term is
computed in skew-symmetric form. This form is known to be less prone to aliasing errors and has better energy-conservation characteristics Kravchenko & Moin
(1997). The solver is based a on a classic two-step pressure-correction scheme with
a third-order backward dierence formula (BDF3) outlined in Guermond et al.
(2006), with boundary conditions used in the Zang-Hussaini algorithm outlined
in Canuto et al. (2007) that should lead to O(t3 ) accuracy. The ltered NavierStokes equations are integrated in time using a combined non-linear viscous step

uj
ui
+
xj
xi

X
1 J1

q uinq
0 ui
t
xj
q=0

J1
X

q Ninq

(3.32)

q=0

with boundary conditions


u |y=0

"

t p n p n1
=
2

0
x
x

(3.33)

y=0

D 2 u |y=Y = 0
t 2 n
D v |y=0
0
t 2 n
= V (x) +
D v |y=Y
0
#
"
t p n p n1
2

=
0
z
z
y=0

(3.34)

v |y=0 =

(3.35)

v |y=Y

(3.36)

w |y=0

D 2 w |y=Y = 0

(3.37)
(3.38)

where 0 , q , q are backward dierence formula of order J coecients, and where


1 ui
1
Ni = uj
+
( uiuj + ijsgs ) Fi .
2 xj
xj 2

(3.39)
37

The resulting Helmholtz equations for ui are solved in spectral space. A pressurecorrection step

p n+1
0  n+1

u
ui +
=0
t i
xi
ui
=0
xi

(3.40)
(3.41)

is then performed to obtain divergence free elds. To avoid applying explicit


Neumann boundary conditions to the pressure, the continuity equation is used to
derive a Poisson equation for vn+1 in spectral space that is equivalent to eq. (3.18),
and subsequently obtaining the pressure and horizontal velocities algebraically
as in equations (3.23), (3.24), and (3.25). The resulting velocity elds are then
divergence free in the entire computational domain including any active sponge
5

regions and satisfy the boundary conditions to O(t 2 ) Guermond et al. (2006).
This time integration scheme is stiy-stable for higher CFL numbers that the
previous vorticity-based scheme Guermond et al. (2006). This property was only
fully utilized in the initialization phase of simulations, whereas the time step was
kept to the same order of magnitude as the vorticity-form spectral code during the
time-averaging phase to ensure direct comparisons could be made.

3.4

Validation

The CTR code was validated against analytical predictions for the growth rate
of the inviscid mixing layer instability, the growth rate of Tollmien-Schlichting (TS) waves in a non-parallel boundary layer, the level of distortion in representing
a convecting Taylor vortex in time, and the pressure uctuations due to sound
scattered by a circular cylinder by Nagarajan (2004). It has been used in a number

38

of DNS and LES of bypass transition in Nagarajan et al. (2007) which reported
good agreement with experiments.
The pseudo-spectral incompressible code in vorticity form was previously validated against linear stability theory for parallel boundary layers (ubl
i = {U(z), 0, 0}
and ubc
i = 0) where errors of less 2% were reported Domaradzki & Metcalfe (1987)
for the T-S wavelengths tested. This validation test was performed again after
modications were made to parallelize the code and replace obscure fast Fourier
transform and linear algebra subroutines with call to standard open source libraries
like LAPACK and FFTW. Results are shown in table 3.1 were consistent and
obtained numerical growth rates matching growth rates predicted by linear stability theory within 5% for a wide range of Reynolds numbers, domain sizes and T-S
wavelengths, discarding 4 outliers (out of 25 tests) where the growth rate predicted
was too close to zero. The same validation test was also performed for the spectral
code in skew-symmetric form, which performed slightly better, with errors less
than 2% with the exception of a few outliers as shown in table 3.2.

39

Table 3.1: Numerical (h E


i) vs theoretical (i ) growth rates of most unstable
t
Orr-Sommerfeld modes in vorticity-form spectral code.
Reynolds number
1805.00
1805.00
1805.00
1805.00
1805.00
2555.00
2555.00
2555.00
2555.00
2555.00
3305.00
3305.00
3305.00
3305.00
3305.00
4055.00
4055.00
4055.00
4055.00
4055.00
4805.00
4805.00
4805.00
4805.00
4805.00

0.500000
0.675000
0.850000
1.02500
1.20000
0.500000
0.675000
0.850000
1.02500
1.20000
0.500000
0.675000
0.850000
1.02500
1.20000
0.500000
0.675000
0.850000
1.02500
1.20000
0.500000
0.675000
0.850000
1.02500
1.20000

i
-0.115938E-01
-0.658913E-02
-0.226416E-02
-0.159557E-03
-0.108623E-02
-0.485721E-02
-0.970083E-03
0.185558E-02
0.267053E-02
0.920218E-03
-0.211583E-02
0.921599E-03
0.283207E-02
0.293234E-02
0.760838E-03
-0.793901E-03
0.162942E-02
0.294678E-02
0.261108E-02
0.192827E-03
-0.915698E-04
0.188049E-02
0.279080E-02
0.216841E-02
-0.414861E-03

i
h % error i
h E
t
-0.115268E-01
0.577699
-0.653586E-02
0.808385
-0.223016E-02
1.50171
-0.154442E-03
3.20600
-0.111583E-02
2.72447
-0.481590E-02
0.850413
-0.935548E-03
3.56002
0.187327E-02
0.953377
0.266628E-02
0.158898
0.888659E-03
3.42944
-0.208443E-02
1.48432
0.945822E-03
2.62837
0.284275E-02
0.377046
0.292242E-02
0.338410
0.729025E-03
4.18128
-0.769478E-03
3.07645
0.164762E-02
1.11693
0.295130E-02
0.153382
0.259848E-02
0.482493
0.164394E-03
14.7455
-0.715104E-04
21.9061
0.189410E-02
0.723902
0.279260E-02 0.645979E-01
0.215651E-02
0.548716
-0.435599E-03
4.99871

40

-1.5
Jones et al.
UDNS
Dyn. Smag.
WALE

-1

Cp

-0.5

0.5

1
0

0.2

0.4

0.6

0.8

(a) NACA0012 airfoil at 5 degree angle of attack at Rec = 0.5 105 Castiglioni et al.
(2014).
0.5
0.45
0.4
0.35

0.3
0.25
0.2
0.15
0.1
0.05
0

1.5

2.5

3.5

4.5

5.5

(b) Flat plate with suction from the top applied at Rex = 105 .

Figure 3.1: Time-average coecients of pressure for airfoils is approximated using


suction boundary conditions for a at plate.

41

0.3
1
y

0.2
0.5
0.1
0

4
x

0.25

0.25

0.2

0.2

0.2

0.2

0.15

0.15

0.15

0.15

0.25

0.25

(a) Contour plot of mean streamwise velocity.

0.1

0.1

0.1

0.1

0.05

0.05

0.05

0.05

0.5
<u>(x=2)

0.5
<u>(x=3)

0.5
<u>(x=4)

0.5
<u>(x=6)

(b) Profiles of mean streamwise velocity before separation (x = 2), after separation
(x = 3), at the recirculating vortex (x = 4), and in the turbulent boundary layer (x = 6).

Figure 3.2: The anatomy of a typical laminar separation bubble over a at plate
taken from large eddy simulation (LES) results using the truncated Navier-Stokes
approach at 3% of DNS resolution.

42

Figure 3.3: Instantaneous iso-surfaces of spanwise vorticity on a at plate due to


suction from the top at Rex = 105 from spectral LES results with the -model at
3% of Spalart & Strelets (2000) DNS resolution.

Figure 3.4: Physical domain, boundary and inlet conditions used to investigate
laminar separation bubble ow over a at plate.

43

(a) Upwind scheme solution

(b) Crank-Nicolson scheme solution

velocity

Figure 3.5: Solutions for the convection equation t T + ux T = 0 after 40 time


steps with CFL=0.8 reproduce from Fletcher (1991)

0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0.1
0.2
0.3
0.4
0.5
0.6

10

Figure 3.6: Spectral LES boundary conditions. Line: top vertical velocity v(x, y =
Y ); dashed line: top streamwise velocity minus unity u(x, y = Y ) 1; dash-dotted
line: wall vertical velocity v(x, y = 0).

44

Table 3.2: Numerical (h E


i) vs theoretical (i ) growth rates of most unstable
t
Orr-Sommerfeld modes in skew-symmetric spectral code.
Reynolds number
1805.00
1805.00
1805.00
1805.00
1805.00
2555.00
2555.00
2555.00
2555.00
2555.00
3305.00
3305.00
3305.00
3305.00
3305.00
4055.00
4055.00
4055.00
4055.00
4055.00
4805.00
4805.00
4805.00
4805.00
4805.00

0.500000
0.675000
0.850000
1.02500
1.20000
0.500000
0.675000
0.850000
1.02500
1.20000
0.500000
0.675000
0.850000
1.02500
1.20000
0.500000
0.675000
0.850000
1.02500
1.20000
0.500000
0.675000
0.850000
1.02500
1.20000

i
-0.115938E-01
-0.658913E-02
-0.226416E-02
-0.159557E-03
-0.108623E-02
-0.485721E-02
-0.970083E-03
0.185558E-02
0.267053E-02
0.920218E-03
-0.211583E-02
0.921599E-03
0.283207E-02
0.293234E-02
0.760838E-03
-0.793901E-03
0.162942E-02
0.294678E-02
0.261108E-02
0.192827E-03
-0.915698E-04
0.188049E-02
0.279080E-02
0.216841E-02
-0.414861E-03

i
h E
t
-0.115823E-01
-0.658011E-02
-0.225150E-02
-0.143196E-03
-0.106467E-02
-0.485301E-02
-0.963559E-03
0.186313E-02
0.268117E-02
0.933770E-03
-0.211210E-02
0.925912E-03
0.283767E-02
0.293782E-02
0.768801E-03
-0.791426E-03
0.163252E-02
0.294896E-02
0.261351E-02
0.199857E-03
-0.896177E-04
0.188215E-02
0.279187E-02
0.217091E-02
-0.403731E-03

h % error i
0.994404E-01
0.136815
0.558928
10.2543
1.98480
0.864110E-01
0.672520
0.406872
0.398365
1.47269
0.176624
0.467992
0.197892
0.186724
1.04665
0.311795
0.190116
0.737562E-01
0.930939E-01
3.64557
2.13182
0.883248E-01
0.382952E-01
0.115136
2.68299

45

Chapter 4
Center for Turbulence Research
Results
Results for three cases computed using the CTR code are reported here: a
benchmark DNS case (DNS), a wall-resolved LES with the dynamic Smagorinsky model (LES), and an under-resolved DNS. Parameters for these simulations
are summarized in Table 4.1. Both the DNS and LES were set up and run by
a collaborator at CTR, Dr Taraneh Sayadi, whereas I set up and performed the
UDNS cases. The DNS by Spalart & Strelet (2000) was initially intended to be
the benchmark case. However, it was run using an incompressible spectral code
with an imposed vorticity-free boundary condition at the top boundary. These top
boundary conditions could not be matched exactly in simulations with the CTR

Nx
Ny
Nz
Ntotal 106
% of spectral DNS
% of CTR DNS
x+
y + at X = 7Y
z +
Sef f ective

Spectral DNS CTR DNS CTR LES CTR UDNS


1022
1536
512
240
120
300
140
90
120
128
32
32
14.7
59.0
2.3
0.7
100
401
15.6
4.7
25
100
3.9
1.2
20
9.7
26.4
57.0
1
0.5
1.0
1.6
6.7
7.6
27.5
29.6
0.25
0.21
0.21
0.20

Table 4.1: Resolution and parameters for all cases run with the CTR code compared to the spectral DNS by Spalart & Strelets (2000).

46

code due to the fringe layer formulation used. The eective top boundary condition is compared with the spectral DNS boundary condition of Spalart & Strelets
(2000) in gure 4.1; the nominal deceleration parameter Sef f ective = 0.21 is less
than for the spectral DNS case (see Table 4.1). Results for additional simulations
performed with a dierent numerical code are reported in Cadieux et al. (2012).
For those additional cases, however, only LES and under-resolved DNS were performed because direct comparisons to the benchmark DNS by Spalart & Strelets
(2000) were made.
CTR DNS, LES, and UDNS were run until the separation bubble stabilized
and turbulent ow was well established downstream of reattachment as illustrated
in gure 4.2 and 4.3. Results were then averaged over multiple bubble breathing
periods. All time-averaged results relating to pressure and friction coecients
obtained are in good qualitative agreement with the DNS benchmark (see gure4.4a and 4.4b).
The wall pressure coecients Cpw = (Pw P )/( 12 U02 ) shown in gure 4.4a
for the UDNS and LES cases are both in good quantitative agreement with the
DNS benchmark with the exception of a slight dierence in bubble length. The
downward slope in Cp in gure 4.4a after x = 5 indicates the existence of a slight
favorable pressure gradient which extends to the end of the physical domain.
This favorable pressure gradient is caused by blowing at the top boundary
seen in gure 4.1.

This presents a limitation in the applicability of results

to the suction side of airfoils in MAVs and blades in turbo-machinery where


such persistent favorable pressure gradients are seldom encountered downstream
of the separation bubble Jones et al. (2010).

Although weak, the favorable

pressure gradient may also articially improve agreement of LES and UDNS

47

0.7
0.6
Normalized velocity

0.5
0.4
0.3
0.2
0.1
0
0.1
1

4
x

Figure 4.1: Normalized wall-normal velocity top boundary condition (V /U0 at


Y = 1): S & S 2000 Spalart & Strelets (2000) (circles), and UDNS (dashed line).
Normalized mean streamwise dierence from freestream velocity ((U U0 )/U0 at
Y = 1): UDNS (line).
results with the DNS benchmark because of its eect on the reattachment location.

At resolutions on the order of 1% of their respective benchmark DNS, and even


without models, all simulations predict the separation point seen in DNS benchmarks exactly. This can be observed in the rst zero-crossing on the wall skin
friction Cf = U
| /( 12 U02 ) plots in gure 4.4b. The UDNS predicts the same
y y=0
shape and maximum value of the peak negative skin friction as the benchmark
DNS. Wall-resolved LES with dynamic Smagorinsky modeling performs slightly
worse than the UDNS run, but still reaches within 15% of the DNS peak negative
skin friction coecient value. UDNS and LES predict the location of the reattachment point with less than 5% dierence with the DNS. UDNS recovers benchmark
48

Figure 4.2: DNS snapshot of iso-surfaces of vorticity: Kelvin-Helmholtz rolls


are visible over the separated shear layer leading to transition to turbulence and
subsequent turbulent ow reattachment, closing of the separation bubble.
DNS results almost exactly for the turbulent Cf in the region downstream of the
bubble whereas LES results underpredict the skin friction in that region.

4.1

Numerical Dissipation

The good quantitative agreement between the no-model highly under-resolved


DNS and benchmark DNS results suggests that the code used may belong to a
category of implicit LES (ILES) where the numerical dissipation plays the role of
subgrid-scale models. As is evident in the results presented in gures 4.4a and 4.4b,
the addition of a subgrid-scale model, even when coupled with higher resolution,
visibly worsens agreement with the DNS benchmark compared to the no-model
case. Such behavior is expected for codes that already provide enough dissipation
49

0.4
y

1
0.2
0

0.5
0
1

4
x

Figure 4.3: Contour plot of normalized average streamwise velocity U/U0 from
the UDNS case. Notice the laminar boundary layer growth followed by a clear
separation bubble spanning from x = 2.8 to x 4.6.
through their numerics so that additional explicit subgrid-scale dissipation is not
required.
The code used has two primary sources of numerical dissipation: truncation
error in derivative approximations, and explicit ltering. Since the code uses sixth
order compact nite dierences and is claimed to be conservative Nagarajan et al.
(2003), focus is placed on quantifying the amount of eective numerical viscosity
introduced by the explicit ltering at each time step Nagarajan (2004). The explicit
high wave number ltering is based on the formulation of a sixth order compact
lter Lele (1992). It is used to remove spurious and unstable high frequency
oscillations that may develop at the interface of the implicitly and explicitly treated
regions due to the codes use of high order nite dierences Nagarajan (2004). This
is done to stabilize the code, and to ensure the implicit and explicit grid solutions
match at their interface. It replaces the use of articial viscosity or newer weighted
essentially non-oscillatory (WENO) type schemes, as well as penalty-type methods
used to stabilize physical or numerical interfaces. Numerical dissipation from this
ltering operation is quantied using two dierent methods.

50

0.5
0.45
0.4
0.35

0.3
0.25
0.2
0.15
0.1
0.05
0

4
x

(a) Coefficient of pressure at the wall.


6

x 10

5
4

Cf

3
2
1
0
1
2
1

4
x

(b) Wall coefficient of friction.

Figure 4.4: Time-averaged Cp and Cf . CTR DNS (circles), CTR LES with
dynamic Smagorinsky model (line), and CTR UDNS (dashed line).

51

4.1.1

Estimating Numerical Dissipation due to Filtering

The amount of eective viscosity the ltering operation imparts to the simulation is estimated by comparing the energy decay rates of runs with ltering
and without ltering. The number of time steps in such an analysis is limited
to 10 to ensure that no numerical instabilities develop. First, two runs are performed using the same value of molecular viscosity, one with ltering and the
other without. Second, the run without ltering is then repeated several times
with larger values of the molecular viscosity until its energy decay curve matches
that of the ltered case.

The excess of the molecular viscosity in a run for

which the best match is achieved provides an estimate of the eective viscosity
that can be attributed to the ltering operation Diamessis et al. (2008). Since
this approach was developed and validated for wakes and isotropic turbulence by
Domaradzki et al. (2003); Domaradzki & Radhakrishnan (2005); Bogey & Bailly
(2006); Diamessis et al. (2008), the procedure is carried through for a limited portion of the domain where a turbulent boundary layer is established after the separation bubble. Incidentally, this is also where the dynamic Smagorinsky model
should be particularly active and where LES predictions were the poorest match
to DNS benchmark data. To conrm this, the same numerical procedure was performed with subgrid-scale modeling turned on. Average eddy viscosities calculated
in the code were output to compare against assessed numerical viscosity.
Total energy curves were obtained for the domain x = 5 to x = 7.5 and y = 0
to y = 0.5 from a set of restart les starting after 30,000 time steps. At such
time, a statistically steady laminar separation bubble ow is well established. A
typical total energy decay curve can be seen in gure 4.5a. Similar results were
obtained using the same procedure with dierent starting iterations (e.g. 36, 000,
and 40, 000). Since the domain selected for the analysis is only a portion of the
52

221,641.1

Total Energy

221,640.9

221,640.7

221,640.5

4
6
Iteration n, where n

total

8
= 30000 + n

10

(a) Total energy decay case


221,644.5
221,644.4
221,644.3

Total Energy

221,644.2
221,644.1
221,644
221,643.9
221,643.8
221,643.7
221,643.6
0

4
6
8
Iteration n, where ntotal = 36000 + n

10

(b) Total energy growth case

Figure 4.5: Total energy in turbulent boundary layer following the laminar separation bubble as a function of time. UDNS (squares), UDNS without ltering (line),
UDNS without ltering and 18% larger molecular viscosity (dashed line), UDNS
without ltering and 33% larger molecular viscosity (dash-dotted line).
53

entire computational domain and the solution is at a statistically steady state, the
energy can either decay or grow over our short analysis times of a few time steps.
However, the described procedure for estimating the eective viscosity is applicable
also to cases where the total energy increases (see gure 4.5b). An over-arching
conclusion from all cases analyzed is that the eect of the ltering operation in the
turbulent boundary layer downstream of reattachment is readily comparable to
increasing the molecular viscosity by 18%. Given the non-dimensional molecular
viscosity in the simulations is 2 106 , the non-dimensional numerical viscosity
is estimated at 0.35 106. The average non-dimensional eddy viscosity in the
same domain calculated using the dynamic Smagorinsky model from the same
starting iteration (30, 000) as in gure 4.5a was 0.71 106 . This corresponds
to approximately 36% of molecular viscosity. The ltering operations numerical
viscosity is thus of the same order of magnitude as the turbulent eddy viscosity
provided by the dynamic Smagorinsky model, making it dicult to distinguish
between eects of the model and that of numerical dissipation.

4.1.2

Quantifying Numerical Dissipation

The amount of eective numerical viscosity is quantied more accurately using


a recently developed technique Schranner et al. (2015). An equation for the evolution of kinetic energy is obtained by multiplying the Navier-Stokes equations by ui .
The residual En of the kinetic energy equation is computed over a given subdomain
from consecutive snapshots of the velocity elds at times tn1 , tn , tn+1 .
En =

Ekin
dV +
t

p
ij
Ekin uj
+ ui
+ ui
xj
xi
xj

dV

(4.1)

54

where
1
Ekin = ui ui
2
2 uk
ui uj
+

ij
ij =
xj
xi
3 xk

Ekin
1 
(Ekin )n+1 (Ekin )n1

t
2t
!
!n
Ekin uj
Ekin uj
p
ij
p
ij

+ ui
+ ui
+ ui
+ ui
xj
xi
xj
xj
xi
xj HO

(4.2)
(4.3)
(4.4)
(4.5)

Eqn (4.1) represents the residual of the kinetic energy transport equation as the
volumetric dissipation rate due to all numerical approximations made in the code
Schranner et al. (2015). It is computed by comparing the rate of change of kinetic
energy using a central dierence formula (4.4) where the kinetic energy is computed
directly from the velocity elds at time tn1 and tn+1 output from the Navier-Stokes
solver used in this case the sixth-order nite dierence CTR code. The exact
transport terms are approximated using the velocity elds at time tn using the
highest order discretization available, denoted by the subscript HO in eqn (4.5).
The residual computed this way can readily be compared to the physical dissipation
rate function
E =

ij

ui
dV
xj

(4.6)

computed at time tn using the same high order discretization, from which an
eective relative numerical viscosity
n
En
=

(4.7)

55

can be calculated provided the physical dissipation rate does not tend to zero. This
analysis is carried out on the same 10 snapshots the previous estimates were computed from for direct comparison. The goal is to provide a more accurate estimate
of numerical viscosity and better explain the role of ltering in this simulation by
repeating the procedure for the 10 snapshots where there is no ltering active.
The numerical dissipation rate and eective numerical viscosity of the CTR
UDNS is computed using nite dierences in the vertical and Fourier expansions
in the horizontal for four regions: in the laminar boundary layer before separation
for x [1.5, 3.0], inside the separated region for x [3.0, 4.5] immediately after
reattachment for x [4.5, 5.5], and further aft of the bubble in the turbulent
boundary layer for x [5.5, 7.0]. These subdomains are also limited in the vertical
to the maximum 99% boundary layer thickness in that region. These quantities
along with the physical dissipation rate for comparison are shown in table 4.2 and
4.3.
As expected, the eective numerical viscosity is nearly zero where the ow is
laminar and well resolved - the numerical dissipation rate is much smaller than
the physical dissipation rate due to the presence of the shear layer. The results
for the ltered case and the unltered case are consistent in the laminar and
separated regions, giving less than 4% dierence. The lter can be said to have no
eect on these regions, and the contribution of the truncation error of the scheme
makes up most of the eective viscosity. In the region near the reattachment
point, the ow experiences strong instabilities and transitions to turbulence. The
Kelvin-Helmholtz rolls and secondary instabilities create sharp streamwise and
vertical gradients. As the shear layer breaks down to turbulence, a number of
small eddies are also created as shown in gure 3.3 and 4.2. These features form
the basis of high wavenumber content which is then promptly removed by the
56

region
laminar
separated
transitional
turbulent

< E > 107 < En > 107 < n / > (%)


8.40
1.20
5.98
4.94
3.30
16.49
10.52
20.66
103.30
21.43
6.38
31.91

Table 4.2: Time-averaged physical and numerical dissipation in 10 CTR UDNS


snapshots.

lter, generating an eective numerical viscosity as large as physical viscosity.


In the turbulent region, all velocities oscillate randomly in space and time at
the smallest scale resolved by the mesh. However, the vertical gradients are now
better resolved due to the larger boundary layer thickness and smaller vertical
uctuations. Filtering explicitly damps the small scales in the streamwise and
spanwise directions, removing energy from the simulation in the turbulent region.
The numerical dissipation rate reaches 31.91% of the physical dissipation rate in
the turbulent boundary layer. This conrms previous estimates that placed the
eective numerical viscosity in the turbulent boundary layer on the same order of
magnitude as the molecular viscosity.
For comparison, the average sgs given by the dynamic Smagorinsky model
in the turbulent region, where it should be at its maximum, is 36% of molecular
viscosity . The numerical dissipation inherent in the CTR code due to ltering
is thus approximately equal to the eddy viscosity predicted by the model. Adding
explicit subgrid-scale modeling to the explicitly ltered simulation thus produces
excessive dissipation, explaining why LES consistently predicts bubble length, peak
negative skin friction, and turbulent skin friction less accurately than UDNS for
this problem.
When comparing the no-lter case to the ltered case, it becomes obvious that
the lter provides the most dissipation in the transitional and turbulent regions,
57

region
laminar (no lter)
separated (no lter)
transitional (no lter)
turbulent (no lter)

< E > 107 < En > 107 < n / > (%)


8.40
1.18
5.90
4.95
2.53
12.63
10.60
2.44
12.19
12.46
0.05
0.27

Table 4.3: Time-averaged physical and numerical dissipation in 10 CTR UDNS


snapshots with ltering turned o.

which fortuitously corresponds to where small scales are expected to be generated


by the physics of the problem, and the same regions where a subgrid-scale model
would be active. The no-lter case has nearly zero eective numerical viscosity
in the turbulent region in comparison to the laminar and separated regions. This
can be explained partly by the absence of sharp vertical gradients observed in
the laminar and separated shear layer that may be under-resolved by the nite
dierence scheme. It could also indicate that the code is dispersive or nearly
unstable in situations where high wavenumber content is present at the numerical
interface between the explicit and implicit grids and nothing is done to smooth
any mismatch.
By ltering high wave number components in all directions regardless of
whether they develop as a result of the physics or the numerics, the explicit lter
eectively removes energy from the simulation and acts as a small-scale energy dissipation mechanism. This ltering operation can thus be regarded as an implicit
subgrid-scale model. For this laminar separation bubble ow and particular simulation parameters, it performed well and produced numerical viscosity comparable
to the average eddy viscosity prescribed by the dynamic Smagorinsky subgrid-scale
model in the correct regions. However, unlike an explicit subgrid-scale model, the
results are unlikely to be repeatable with a dierent code or a dierent grid.

58

Chapter 5
Spectral Results I
The skew-symmetric spectral large eddy simulation tool lsbflow detailed in
section 3.3.3 is used to unambiguously assess the performance of dierent subgridscale models. The following simulations of a laminar separation bubble over a at
plate were performed at 3% and 1% of the corresponding DNS benchmark resolution: a no-model LES or under-resolved DNS to serve as a baseline (UDNS),
an LES with the dynamic Smagorinsky model (DSM), another with the -model
(SIG), and one using the truncated Navier-Stokes approach with automatic ltering (TNS). Each simulation is started from the same initial condition before
the ow has transitioned to turbulence. All simulations were run until the separation bubble stabilized and turbulent ow was well established downstream of
reattachment before starting to collect statistics. Time-averaging was done over 40
non-dimensional time units t = t LUx0 , well beyond the point of reaching averagingperiod independence. Parameters for these simulations are summarized in Table
5.1.
Results obtained sharply contrast previous LES results using sixth-order nite
dierences with explicit ltering where the no-model simulation provided the best
agreement and adding the dynamic Smagorinsky model deteriorated agreement
with the benchmark DNS Cadieux et al. (2014). Spectral under-resolved DNS
results instead demonstrate poor agreement with the benchmark DNS at 3% of
DNS resolution and even lose the ability to predict the separation point at 1%.
Such disagreement is expected because the smallest relevant physical scales of the
59

Table 5.1:
Resolution and parameters for LES cases and DNS by
Spalart & Strelets (2000).
Nx
Ny
Nz
Ntotal 106
% of DNS
x+ at x = 7
y + at x = 7
z + at x = 7

DNS 3% LES 1% LES


1022
256
240
120
56
32
120
32
24
14.7
0.46
0.18
100
3.1
1.25
12
54
54
1
0.2
0.27
6.3
26
34

laminar separation bubble problem are not captured with such coarse grids and the
incompressible spectral code provides little dissipation due to approximations in
the discretization of the Navier-Stokes equations in space and time. The no-model
simulations instead suer from a known problem common to all poorly resolved
spectral simulations: slow accumulation of energy in the largest resolved wave
numbers. This can in turn cause numerical oscillations and articial backscatter
which contaminates the larger scales, creating a feedback loop. Due to this phenomenon, UDNS predictions at 1% are no longer representative of the physics of
the problem.
The addition of a subgrid-scale model to provide dissipation successfully stops
energy from accumulating in the smallest resolved scales and markedly improves
results in all cases. Explicitly-modeled LES capture separation, transition, and
turbulent reattachment accurately on both grids. The instantaneous spanwise
vorticity and vertical streamwise velocity gradient for the truncated Navier-Stokes
simulation at 3% of the benchmark DNS resolution are presented in gures 5.1a and
5.1b to illustrate the laminar, transitional, and turbulent regions and the scale of
their structures. The capability to predict accurately at low computational cost the

60

0.4
y

0.3
0.2
0.1
0

4
x

(a) Contours of spanwise vorticity for a slice along the x-y plane at z = 0.
0.6

0.4
0.2
0

4
x

(b) Contours of streamwise velocity vertical gradient at the wall

u
y |y=0 .

Figure 5.1: Instantaneous LES results with TNS model at 3% of DNS resolution
showing shear layer break up and ow reattachment.
2
2
average skin friction Cf = 2Uy /(U
), pressure coecient Cp = 2(pp )/(U
),

and the location of separation and reattachment is of particular interest to airfoil and blade designers. The performance of the LES performed is evaluated
using the above-mentioned quantities by comparing to the spectral DNS data by
Spalart & Strelets (2000). Other quantities such as boundary layer thicknesses,
mean and RMS velocity proles are also investigated to further characterize the
performance of the dierent subgrid-scale models.

5.1

LES at 3% of DNS Resolution

The coecient of skin friction and coecient of pressure predictions from each
simulation is displayed in gures 5.2a and 5.2b along with the benchmark DNS data

61

by Spalart & Strelets (2000). All 3% LES cases capture the location of the separation point exactly, visible in the rst zero crossing of the skin friction coecient Cf .
Beyond the peak negative friction location, UDNS results strongly depart from the
benchmark DNS. A larger separation bubble is observed. The delayed rise in the
coecient of pressure, as well as the skin friction going from negative to positive
indicate a later reattachment point. The skin friction downstream of reattachment
is also poorly predicted by the UDNS. In contrast, the simulations with an explicit
subgrid-scale model recover both Cp and Cf curves of the benchmark DNS almost
exactly. A small delay in the onset of the negative peak skin friction is observed
in all models. Similarly, a small discrepancy in nominal pressure in the turbulent
boundary layer downstream of reattachment is observed. These discrepancies are
likely related to the rapid coarsening of the grid due to the algebraic mapping.
Figure 5.3 displays similar agreement for all LES with explicit models for the
boundary layer thicknesses calculated using a pseudo-velocity Up based on the
vertical integral of spanwise vorticity as follows
Up (x, y) =

v(x, y ) u(x, y )
dy .

x
y

(5.1)

The displacement thickness is slightly over-predicted in the separated region, which


could be due in part to the rapid coarsening of the grid at that vertical location.
It is very dicult to distinguish between models in LES at 3% of DNS resolution,
each recovering the pressure, skin friction and boundary layer thicknesses very
accurately, discounting small grid-related discrepancies.
Mean velocity proles shown in gure 5.4 show minor disagreements with the
benchmark DNS in terms of the vertical location of the separated shear layer as
well as the amount of reverse ow present at x = 3.5. RMS uctuation velocities

62

are found in good agreement for the dynamic Smagorinsky model and -model,
whereas the truncated Navier-Stokes predictions agree perfectly at x = 2.5 but
depart slightly from the benchmark at x = 3.5.

5.2

LES at 1% of DNS Resolution

Further reducing the number of collocation points in each direction to reach 1%


of the DNS resolution changes the performance of LES with explicit models only
slightly. In contrast, the no-model simulation now converges to a ow eld where
the ow separates later, but reattaches and transitions quickly to turbulence without forming a proper laminar separation bubble. This is made evident in gures
5.5b and 5.5a by the later rst Cf zero crossing of the UDNS, its earlier second
zero crossing, and its Cp curves smooth rise without a characteristic plateau.
Explicitly-modeled LES show consistent improvement in agreement with the
benchmark DNS when compared to the UDNS for each key feature of the ow as
shown in table 5.2 and in gures 5.5b and 5.5a. This improvement is directly linked
with their subgrid-scale dissipation contribution ijsgs Sij . This is shown in gure
5.6 where the models subgrid-scale dissipation is plotted. Notice that the strongest
contributions for all models are near x = 3.5 where the separated shear layer breaks
up as is visualized in gure 3.3 using iso-surfaces of spanwise vorticity. The eective
turbulent eddy viscosity reaches sgs 10 there for both dynamic Smagorinsky
and models. sgs then decreases until it is O() at reattachment, and further
decreases downstream in the turbulent boundary layer. As expected, the models
are not active in the laminar and freestream region, but do have some impact inside

63

0.5
0.45
0.4
0.35

0.3
0.25
0.2
0.15
0.1
0.05
0
1

4
x

(a) Pressure coefficients.


3

x 10

5
4
3

Cf

2
1
0
1
2
3
4

4
x

(b) Skin-friction coefficients.

Figure 5.2: Pressure and friction coecients for LES at 3% of DNS resolution.
Circles: Spalart & Strelets (2000) DNS; diamonds: under-resolved DNS; dashed
line: -model; dash-dotted line: dynamic Smagorinsky; line: truncated NavierStokes.
64

0.25

0.2

0.15

0.1

0.05

4
x

(a) Displacement thickness .


0.07
0.06
0.05

0.04
0.03
0.02
0.01
0

4
x

(b) Momentum thickness .

Figure 5.3: Boundary layer thicknesses for LES results at 3% of DNS resolution. Squares: Blasius solution; Circles: Spalart & Strelets (2000) DNS; diamonds:
under-resolved DNS; dashed line: -model; dash-dotted line: dynamic Smagorinsky; line: truncated Navier-Stokes.
65

0.2

0.18

0.18

0.16

0.16

0.14

0.14

0.12

0.12

0.1

0.1

0.2

0.08

0.08

0.06

0.06

0.04

0.04

0.02

0.02

0.5

1.5

2.5

0.2

0.2

0.18

0.18

0.16

0.16

0.14

0.14

0.12

0.12

0.1

0.1

0.08

0.08

0.06

0.06

0.04

0.04

0.02

0.02

0.5

1.5

(c) -model.

1.5

2.5

(b) dynamic Smagorinsky.

(a) under-resolved DNS.

0.5

2.5

0.5

1.5

2.5

(d) truncated Navier-Stokes.

Figure 5.4: Proles in LES at 3% of DNS resolution. Line: U; dashed line:


u /umax . From left to right, plots spaced by 1.5: x = 2.5, x = 3.5.

66

the separated shear layer as early as x = 2.5, corresponding to the earliest nonzero turbulent uctuations in v as plotted in gure 5.10b. The average subgridscale dissipation provided by the truncated Navier-Stokes approach is estimated
by summing the energy removed each time the automatic ltering operation is
triggered as denoted by tm over one non-dimensional time unit t = t LUx0 = Nt t
for which the lter is triggered M Nt /200 times:
hijsgs Sij i

M h
i
1 X
E E
t=tm
Nt t m=1

"

3
M
X
1
1 X
(ui ui)(ui ui)
=
Nt t m=1 i=1 2

(5.2)
#

(5.3)
t=tm

This estimate is not representative of the dynamics of the truncated Navier-Stokes


approach, but it would seem to indicate that its average dissipation contribution
is one order of magnitude smaller than the dynamic and models. The maximum
eective viscosity of the truncated Navier-Stokes approach is O() and is located
aft of x = 3.5. Its downstream location and negligible contributions before x = 3
may explain why truncated Navier-Stokes results better capture the onset of the
peak negative skin friction.
Truncated Navier-Stokes predicitons reach remarkable quantitative agreement
with the benchmark DNS for Cf and Cp with an overall RMS error lower than
the other two models. Its only departure is seen from x = 4.5 to x = 5 where it
over-predicts the skin friction just downstream of reattachment by approximately
10%. LES with the -model obtains the second best agreement, followed closely
by the dynamic Smagorinsky model. Both eddy viscosity models do not recover
the turbulent skin friction coecient as well as their 3% counterparts, especially
inside the bubble from x = 3 to x = 3.5. This is explained in part by the more
rapidly coarsening grid in the vertical, but also because of their dissipative eect
67

on all scales of motion (as opposed to only the smallest resolved scales). The
dynamic Smagorinsky also under-predicts the peak negative skin friction and the
reattachment location.
Displacement and momentum thickness results seen in gures 5.7a and 5.7b for
all explicitly modeled LES are consistent with their corresponding higher resolution
results and show good agreement with the DNS benchmark. A clear departure is
again observed for UDNS. Its displacement thickness conrms the lack of a typical
separation bubble, whereas its momentum thickness indicates a quicker transition
and relaxation to a near-equilibrium turbulent boundary layer. This is further
evidenced in the agreement of the mean UDNS velocity in wall units with the log
law of the wall near the outow at x = 7 shown in gure 5.8. Perfect agreement
with the log law of the wall was not reported in the DNS by Spalart & Strelets
(2000). It is not seen in explicitly-modeled LES. The turbulent boundary layer
they predict has yet to return to equilibrium at x = 7, more in line with the
DNS benchmark. All simulations capture the viscous sublayer accurately. This is
expected as the algebraic mapping is designed to be strong enough to resolve the
inlet Blasius boundary layer and gives us a y + of 0.27 at x = 7.
The performance of the models is further delineated in the mean velocity and
RMS uctuation velocity proles shown in 5.9. RMS uctuation velocities are
computed from 40 ow elds taken approximately one non-dimensional time unit
apart substracted from the mean ow which is averaged over a total of 40 nondimensional time units. The mean and RMS velocity proles are averaged in the
spanwise direction. The departures of the dynamic Smagorinsky and models
seen in the skin friction curve between x = 3 and x = 3.5 are mirrored in their
mean velocity proles and associated vertical derivatives. A stronger reverse ow

68

Table 5.2: Performance of subgrid-scale models at 1% of DNS resolution.


Cp average % error
Peak Cf % error
Reattachment point % error
Turbulent Cf average % error
Reattachment shape factor % error
Rel. CPU cost

UDNS DSM SIG TNS


51
7.8 4.2
1.6
52
17 3.4
3.9
29
4.4 1.4
0.5
2.5
2.9 4.7
5.1
9.1
10 8.4
2.2
1.0
2.5 1.4
1.2

than found in the DNS is predicted by both models, whereas the truncated NavierStokes results compare well with the benchmark. On the other hand, the RMS
uctuation velocity proles of all explicitly modeled LES generally display satisfactory agreement with the benchmark. UDNS proles conrm what is visible in
its Cf curve, that the ow transitions and reattaches between x = 3 and x = 3.5.
The maximum RMS uctuation of the streamwise and vertical velocity as a
function of streamwise distance is shown in gures 5.10 along with the benchmark
DNS data. Only the truncated Navier-Stokes approach recovers accurately the
peak uctuation velocities, whereas all models predict the uctuation intensities
downstream of reattachment beyond x = 5 satisfactorily.
The accuracy with which each subgrid-scale model recovers key features of
the ow of critical importance to airfoil and blade designers such as the pressure
and skin friction, as well as the shape factor at reattachment are quantied by
computing the relative percent error with respect to the benchmark DNS and
shown in table 5.2. The computational cost of each subgrid-scale model is measured
in terms of average number of iterations per second (performed on the same number
of MPI processes on the same machine). It is displayed in table 5.2 as the ratio of
UDNS speed (iterations/second) to that of the modeled simulations to illustrate
the extra computational time required for each model. The computational cost

69

0.5
0.45
0.4
0.35

0.3
0.25
0.2
0.15
0.1
0.05
0
1

4
x

(a) Pressure coefficients.


3

x 10

5
4
3

Cf

2
1
0
1
2
3
4

4
x

(b) Skin-friction coefficients.

Figure 5.5: Pressure and skin friction coecients for LES at 1% of DNS resolution.
Circles: Spalart & Strelets (2000) DNS; diamonds: under-resolved DNS; dashed
line: -model; dash-dotted line: dynamic Smagorinsky; line: truncated NavierStokes.
70

of the dynamic procedure in the dynamic Smagorinsky model is higher than that
of nding the singular values of the derivative tensor in the -model and nearly
twice that of the periodic ltering in TNS. This is due to the communication cost
incurred in the multiple test-ltering operations in the dynamic procedure.
It should be noted that the dynamic Smagorinsky models relative computational cost appears higher than previously reported by Piomelli for a comparable
spectral code Piomelli (1999). The dierence is not in model implementation, but
in the algorithm chosen for the non-linear step. A three-step Runge-Kutta scheme
(RK3) requires three times as many Fourier transforms as the scheme used here,
and thus takes roughly three times as long to compute. Piomellis implementation
uses RK3 but performs the dynamic procedure only once Piomelli (1999). This is
computationally equivalent to performing the dynamic procedure every third iteration in the solver presented here. It is thus reasonable that the relative cost of the
dynamic Smagorinsky model to the no-model simulation reported here is nearly
three times the one reported by Piomelli. Such dierences are to be expected in
dierent Navier-Stokes solvers because the relative cost corresponds to the ratio
of the total number of operations, including all the extra operations added by the
model, to the number of operations required by the solver alone.
Overall, the truncated Navier-Stokes approach and -model both outperform
dynamic Smagorinsky models due to their better agreement with the DNS benchmark and their lower computational cost. The accuracy of the TNS results at
1% of DNS resolution demonstrates that the use of periodic ltering alone can
provide enough dissipation to act as a substitute for an explicit subgrid-scale
model. It also lends weight to the applicability of the TNS approach beyond
isotropic turbulence and channel ow previously tested Domaradzki et al. (2002);
Tantikul & Domaradzki (2010, 2011). Its better performance is attributed to the
71

ltering operation only aecting the smallest resolved scales, whereas turbulent
eddy viscosity models often have a dissipative eect on a wider range of scales.
Excluding the drastic changes in the baseline no-model simulation (UDNS)
results in going to the coarser mesh, LES results at 3% and 1% are consistent in
their ability to predict Cf , Cp , and boundary layer thicknesses. Consistency of
results for each respective model at these dierent resolutions indicates a healthy
degree of grid independence, which is seldom obtained at such low resolutions.
The model results match those of the dynamic Smagorinsky model but improve
upon it in the separated shear layer where intense shear is present, vindicating its
stated purpose Nicoud et al. (2011).

5.3

Spanwise Structures

To investigate the importance of spanwise structures in laminar separation


bubble ow, the velocity auto-correlation functions were computed with respect to
spanwise distance at dierent locations in the streamwise direction at a height equal
to the displacement thickness for the 1% LES case using the truncated NavierStokes approach with automatic ltering. They were then ensemble-averaged over
35 realizations separated by approximately one non-dimensional time unit. The
plots of these auto-correlation functions in gure 5.11 have a number of interesting
features. gure 5.11a shows that the streamwise velocity uctuations are strongly
correlated from x = 2 to x = 4 where the eects of the suction boundary layer
are strongest, whereas elsewhere in the domain, Ruu tends to zero with distance r.
Similarly, Rvv shows a strong correlation nearest to the suction location at x = 3,
but otherwise tends to zero. Spanwise uctuations Rw w all tend to zero more
rapidly except for locations where the suction is strongest at x = 3 and where the

72

presence of Kelvin-Helmholtz rolls have been observed around x = 4. That the


spanwise uctuations do not tend to zero near the bubble likely indicates one of
two things: a lack of spanwise resolution or that the spanwise structures are larger
than the spanwise domain size.
A truncated Navier-Stokes simulation with automatic ltering where the extent
of the domain and the number of collocation point in the spanwise direction were
doubled was performed to test whether the size of the spanwise structures was
not well captured in previous simulations. The auto-correlation functions are computed and ensemble-averaged in the same way over 35 realizations. The gure 5.12
displays each velocity correlation function at dierent streamwise locations. From
comparison with gure 5.11, it is immediately visible that most velocity autocorrelations are more closely clustered around zero at r = 20 than their counterparts on the smaller spanwise domain. One exception remains across all velocities
at x = 3 where the boundary conditions impose a structure on the ow. Another
exception is seen in the streamwise velocity correlation at x = 2 which is the
location where separation occurs, again driven by the suction boundary condition.
However, previous non-zero correlations at x = 4 are no longer visible, indicating
that the turbulent structures are now smaller than the domain extent. For future
very coarse LES of laminar separation bubble over a at plate, the spanwise extent
should be larger than that used in Spalart & Strelets (2000). Although doubling
the spanwise size may not be necessary, a minimum of Lz = 1Y corresponding
to Lz = 0.5Lb where Lb is the bubble length should be used to ensure velocity
auto-correlations fall to zero.
Resolution also plays an important role in the size of spanwise structures. To
visualize the dierence, the velocity auto-correlation functions were computed and
ensemble-averaged in the same fashion for a truncated Navier-Stokes simulation
73

with automatic ltering at 3% of the DNS resolution. The number of points in the
spanwise direction is increased from the 1% LES-S by a factor of 34 , going from a
z + = 34 to z + = 26. This results in an improvement in the tendency of all
correlations to go to zero with distance r as seen in gure 5.13, but the improvement
is not as dramatic as for a larger spanwise domain size. Since increasing resolution
alone may not be enough, the earlier conclusion that the domain size should be
increased in future coarse LES of laminar separation bubble ows over at plates
stands.

74

3
0.4

3.5

0.3
y

4
0.2
4.5

0.1
0
2

2.5

3.5

4.5

5.5

6.5

7.5

(a) Dynamic Smagorinsky model.


3
0.4

3.5

0.3
y

4
0.2
4.5

0.1
0
2

2.5

3.5

4.5

5.5

6.5

7.5

(b) model.
6.5

0.4

0.2

7.5

0.3

0.1
0
2

8
2.5

3.5

4.5

5.5

6.5

7.5

(c) truncated Navier-Stokes.

Figure 5.6: Contours of the logarithm of time and spanwise-averaged subgrid-scale


dissipation log10 (hijsgs Sij i). Each contour represents a jump of 0.25.

75

0.2
0.18
0.16
0.14

0.12
0.1
0.08
0.06
0.04
0.02
0

4
x

(a) Displacement thickness .


0.07
0.06
0.05

0.04
0.03
0.02
0.01
0

4
x

(b) Momentum thickness .

Figure 5.7: Boundary layer thicknesses for LES at 1% of DNS resolution. Squares:
Blasius solution; Circles: Spalart & Strelets (2000) DNS; diamonds: under-resolved
DNS; dashed line: -model; dash-dotted line: dynamic Smagorinsky; full line:
truncated Navier-Stokes.
76

25
20
u+

15
10
5
0

10

10

10

10

y+

Figure 5.8: Turbulent boundary layer mean streamwise velocity in wall units at
x = 7 for LES at 1% of DNS resolution. Squares: log law of the wall; circles:
u+ = y + ; diamonds: under-resolved DNS; dashed line: -model; dash-dotted line:
dynamic Smagorinsky; full line: truncated Navier-Stokes.

77

0.2

0.18

0.18

0.16

0.16

0.14

0.14

0.12

0.12

0.1

0.1

0.2

0.08

0.08

0.06

0.06

0.04

0.04

0.02

0.02

0.5

1.5

2.5

0.2

0.2

0.18

0.18

0.16

0.16

0.14

0.14

0.12

0.12

0.1

0.1

0.08

0.08

0.06

0.06

0.04

0.04

0.02

0.02

0.5

1.5

(c) -model.

1.5

2.5

(b) dynamic Smagorinsky.

(a) under-resolved DNS.

0.5

2.5

0.5

1.5

2.5

(d) truncated Navier-Stokes.

Figure 5.9: Proles in LES at 1% of DNS resolution (lines), and in DNS (symbols).
Line, circles: U; dashed line, diamonds: u /umax . From left to right, plots spaced
by 1.5: x = 2.5, x = 3.5.

78

0.25

0.2

0.15

0.1

0.05

4
x

(a) umax (x).


0.25

0.2

0.15

0.1

0.05

4
x

(b) vmax
(x).

Figure 5.10: Maximum RMS uctuation velocity in LES at 1% of DNS resolution.


Circles: Spalart & Strelets (2000) DNS; diamonds: under-resolved DNS; dashed
line: -model; dash-dotted line: dynamic Smagorinsky; full line: truncated NavierStokes.
79

1
0.8

uu

0.6
0.4
0.2
0
0.2
0.4
0

0.05

0.1

(a) Ruu =

0.15
r=nz

0.2

0.25

0.3

0.25

0.3

0.25

0.3

<u (x,y,z)u (x,y,z+r)>


<u (x,y,z)u (x,y,z)>

1
0.8

ww

0.6
0.4
0.2
0
0.2
0.4
0

0.05

0.1

(b) Rvv =

0.15
r=nz

0.2

<v (x,y,z)v (x,y,z+r)>


<v (x,y,z)v (x,y,z)>

1
0.8

vv

0.6
0.4
0.2
0
0.2
0.4
0

0.05

0.1

(c) Rww =

0.15
r=nz

0.2

<w (x,y,z)w (x,y,z+r)>


<w (x,y,z)w (x,y,z)>

Figure 5.11: Ensemble-averaged spanwise auto-correlation functions for TNS at


1% of DNS resolution. Blue squares: x = 1; light green circles: x = 2; red crosses:
x = 3; black diamonds: x = 4; green triangles: x = 5; fuschia dash-dotted line:
x = 6; teal dashed line: x = 7.
80

1
0.8

uu

0.6
0.4
0.2
0
0.2
0.4
0

0.1

0.2

(a) Ruu =

0.3
r=n z

0.4

0.5

0.6

0.5

0.6

0.5

0.6

<u (x,y,z)u (x,y,z+r)>


<u (x,y,z)u (x,y,z)>

1
0.8

ww

0.6
0.4
0.2
0
0.2
0.4
0

0.1

0.2

(b) Rvv =

0.3
r=n z

0.4

<v (x,y,z)v (x,y,z+r)>


<v (x,y,z)v (x,y,z)>

1
0.8

vv

0.6
0.4
0.2
0
0.2
0.4
0

0.1

0.2

(c) Rww =

0.3
r=n z

0.4

<w (x,y,z)w (x,y,z+r)>


<w (x,y,z)w (x,y,z)>

Figure 5.12: Ensemble-averaged spanwise auto-correlation functions for TNS with


doubled spanwise domain size and spanwise collocation points. Blue squares: x =
1; light green circles: x = 2; red crosses: x = 3; black diamonds: x = 4; green
triangles: x = 5; fuschia dash-dotted line: x = 6; teal dashed line: x = 7.
81

1
0.8

uu

0.6
0.4
0.2
0
0.2
0.4
0

0.05

0.1

(a) Ruu =

0.15
r=nz

0.2

0.25

0.3

0.25

0.3

0.25

0.3

<u (x,y,z)u (x,y,z+r)>


<u (x,y,z)u (x,y,z)>

1
0.8

ww

0.6
0.4
0.2
0
0.2
0.4
0

0.05

0.1

(b) Rvv =

0.15
r=nz

0.2

<v (x,y,z)v (x,y,z+r)>


<v (x,y,z)v (x,y,z)>

1
0.8

vv

0.6
0.4
0.2
0
0.2
0.4
0

0.05

0.1

(c) Rww =

0.15
r=nz

0.2

<w (x,y,z)w (x,y,z+r)>


<w (x,y,z)w (x,y,z)>

Figure 5.13: Ensemble-averaged spanwise auto-correlation functions for TNS at


3% of DNS resolution. Blue squares: x = 1; light green circles: x = 2; red crosses:
x = 3; black diamonds: x = 4; green triangles: x = 5; fuschia dash-dotted line:
x = 6; teal dashed line: x = 7.
82

Chapter 6
Spectral Results II
Prior to the completion of the development of the spectral large eddy simulation
tool lsbflow, another spectral solver described in section 3.3.2 was used to perform
comparable simulations. LES at 3% of DNS resolution were performed with the
truncated Navier-Stokes approach with a constant ltering period corresponding
to 0.5% of one non-dimensional time unit and compared to a no-model simulation.
LES at 1% of DNS resolution were done with three subgrid-scale models: the
dynamic Smagorinsky model (DSM), the model (SIG), and the truncated NavierStokes approach with a constant ltering period (TNS). An under-resolved DNS
(UDNS) was also performed to serve as a base line comparison. Parameters for
these simulations are summarized in table 6.1 and denoted by LES-V to distinguish
them from previous LES obtained with lsbflow. Simulations were run until the
separation bubble stabilized and turbulent ow was well established downstream
of reattachment. Both 3% and 1% simulation results were time-averaged over more
than 25 non-dimensional time units t = t LUx0 .

6.1

LES-V at 3% of DNS Resolution

At 3% resolution, no-model results display a larger separation bubble and do


not recover the expected skin friction levels downstream of reattachment. This is
consistent with previous spectral results, but the UDNS predictions are in even
greater disagreement with the DNS benchmark in this case as is evidenced in

83

DNS 3% LES-V 1% LES-V


Nx
1022
240
224
Ny
120
56
32
Nz
120
32
24
Ntotal 106
14.7
0.43
0.17
% of DNS
100
2.9
1.1
+
x at x = 7
12
54
58
y + at x = 7
<1
0.07
0.27
z + at x = 7
6.3
26
34
Table 6.1:
Resolution and parameters for LES-V cases and DNS by
Spalart & Strelets (2000).

gures 6.1a, 6.1b, and 6.2. Discrepancies in Cf and Cp around x = 3 seen in


the UDNS could be an indication of the no-model simulation reaching a dierent
steady state with stronger and perhaps slower bubble breathing cycles which
were not smoothed out by the time averaging.
Truncated Navier-Stokes results improve on the no-model simulation drastically. The Cp curve is well predicted with the exception of a later rise in pressure,
indicating a slightly larger bubble. This is conrmed by its downstream second
zero crossing of the Cf curve and its larger displacement thickness in the separated
ow region visibile in gure 6.2a. TNS overpredicts the peak negative skin friction
by 12%, predicts the reattachment point accurately with 2.4% relative error, and
under-predicts the turbulent skin friction downstream of reattachment by 5.3%.
The level of mean reverse ow it predicts at x = 3.5 is in good agreement with
the benchmark DNS as shown in gure 6.3. Departures from the benchmark DNS
in its mean velocity and RMS proles at that location occur where the resolution
coarsens rapidly and the boundary layer thickness is at its largest.
Notice that the y + for the 3% LES is much smaller than for typical wallresolved LES in which y + 1. It is in fact signicantly smaller than required

84

for DNS y + 0.5. This means that many points are clustered very close to
the wall and that very few points are used to resolve vertical locations beyond
y = 0.2, which is where the separated shear layer is found to transition. Given
that previous investigators have identied resolving the spearated shear layer break
up and transition as the key to obtaining agreement, it is both surprising and
encouraging that the truncated Navier-Stokes approach managed to recover the
Cp and Cf curves this accurately. Conversely, this dierence explains most of the
inconsistencies found between the previous 3% LES results and the 3% LES-V
results.

6.2

LES-V at 1% of DNS Resolution

The grid is signicantly coarsened in the vertical direction to reach 1% of


the DNS resolution. The stretching parameter is also changed to ensure a larger
y + which matches the previous 1% LES. These changes gravely deteriorate the
no-model simulation predictions. As seen in gures 6.4a and 6.4b, the UDNS
now predicts a later separation point but a quicker transition to turbulence and
reattachment. This is largely consistent with previous 1% UDNS results. One
exception is that the under-resolved DNS turbulent Cf levels downstream of reattachment are under-predicted by nearly 40%. Another is that its mean velocity
prole in wall units does not agree with the log law of the wall at x = 7 as exemplied in gure 6.6, whereas it did in previous results (see gure 5.8). The reason for
this discrepancy between spectral UDNS results can only be numerical in nature,
because all other parameters were kept constant. In any case, both UDNS suer
from excessive energy accumulation in the small scales which eventually starts to

85

0.5
0.45
0.4
0.35

Cp

0.3
0.25
0.2
0.15
0.1
0.05
0
1

4
x

(a) Pressure coefficients.


3

x 10

Cf

4
x

(b) Skin-friction coefficients.

Figure 6.1: Pressure and friction coecients for LES-V at 3% of DNS resolution.
Circles: Spalart & Strelets (2000) DNS; diamonds: under-resolved DNS; dashed
line: -model; dash-dotted line: dynamic Smagorinsky; line: truncated Navier86
Stokes.

0.25

0.2

0.15

0.1

0.05

4
x

(a) Displacement thickness .


0.07

0.06

0.05

0.04

0.03

0.02

0.01

4
x

(b) Momentum thickness .

Figure 6.2: Boundary layer thicknesses for LES-V results at 3% of DNS resolution. Squares: Blasius solution; Circles: Spalart & Strelets (2000) DNS; diamonds:
under-resolved DNS; dashed line: -model; dash-dotted line: dynamic Smagorin87
sky; line: truncated Navier-Stokes.

0.2

0.18

0.18

0.16

0.16

0.14

0.14

0.12

0.12

0.1

0.1

0.2

0.08

0.08

0.06

0.06

0.04

0.04

0.02

0.02

0.5

1.5

(a) under-resolved DNS.

2.5

0.5

1.5

2.5

(b) truncated Navier-Stokes.

Figure 6.3: Proles in LES-V at 3% of DNS resolution. Line: U; dashed line:


u /umax . From left to right, plots spaced by 1.5: x = 2.5, x = 3.5.

drive motion at the large scales and pollute results in ways that are not representative of the underlying physics.
By contrast, the explicitly-modeled LES predictions show improved agreement
with the benchmark DNS. All models predict the pressure distribution and reattachment location very accurately. The truncated Navier-Stokes approach with
a xed ltering period performs overall better on the coarser grid than it did on
the 3% grid. This is attributed to the better distribution of points away from the
viscous sublayer toward the midplane so as to better resolve the separated shear
layer at its thickest. It is the only model to the recover the turbulent skin friction
downstream of reattachment accurately as seen in gure 6.4b. It has the lowest
88

RMS error for Cf and Cp . However, it overpredicts the peak negative skin friction
as noted in table 6.2 and the size of the separated displacement thickness by 5.9%
as shown in gure 6.5. It shows the best agreement for the mean velocity prole
at x = 3.5 seen in gure 6.7, and for the streamwise distribution of the maximum
RMS velocity umax (x). The dynamic Smagorinsky model displays the second best
agreement with DNS data. It recovers the skin friction, pressure, displacement
and momentum thickness curves of the DNS benchmark almost exactly. However, its predicted skin friction levels are consistently below the DNS curve beyond
the reattachment point. This is reected in the larger maximum RMS velocity
umax (x) aft of x = 4.5 in gure 6.8a. The same behavior is observable in the
model results for Cf and umax (x). Both the dynamic Smagorinsky and models
display stronger reverse ow in gure 6.7 than TNS and the DNS benchmark at
x = 3.5. The model signicantly overshoots the peak negative skin friction. This
is surprising as its subgrid-scale dissipation contribution was previously shown to
be almost identical to the dynamic Smagorinsky models as shown in gure 5.6.
All models otherwise agree fairly well, but not exactly with the log law of the wall
as displayed in gure 6.6, in agreement with Spalart & Strelets (2000) who did not
report perfect agreement.
Results obtained with this vorticity-form solver largely corroborate conclusions
drawn from lsbflow results. Both sets of simulations demonstrate that accurate
LES are possible at 1% of DNS resolution. The only notable discrepancy is the
LES-V models consistent under-prediction of the pressure and skin friction downstream of reattachment. All other parameters having been kept constant, this must
be due to implementation dierences. Namely, the exact way the top boundary
conditions are enforced for the streamwise velocity diers between the two spectral
solvers, and the vorticity-form is known to be more prone to aliasing errors and
89

Table 6.2: Performance of subgrid-scale models in LES-V at 1% of DNS resolution.


Cp average % error
Peak Cf % error
Reattachment point % error
Turbulent Cf average % error
Reattachment shape factor % error

UDNS DSM SIG TNS


46
3.2 3.4
4.0
37
5.6
35
13
20
1.1 2.0
1.0
38
14
19
8.9
9.7
7.3
18
8.3

Gibbs oscillations when highly under-resolved, which could impede the models
ability to predict the correct subgrid-scale dissipation.

90

0.5
0.45
0.4
0.35

0.3
0.25
0.2
0.15
0.1
0.05
0
1

4
x

(a) Pressure coefficients.


3

x 10

5
4
3

Cf

2
1
0
1
2
3
4

4
x

(b) Skin-friction coefficients.

Figure 6.4: Pressure and friction coecients for LES-V at 1% of DNS resolution.
Circles: Spalart & Strelets (2000) DNS; diamonds: under-resolved DNS; dashed
line: -model; dash-dotted line: dynamic Smagorinsky; line: truncated NavierStokes.
91

0.2
0.18
0.16
0.14

0.12
0.1
0.08
0.06
0.04
0.02
0

4
x

(a) Displacement thickness .


0.07

0.06

0.05

0.04

0.03

0.02

0.01

4
x

(b) Momentum thickness .

Figure 6.5: Boundary layer thicknesses for LES-V at 1% of DNS resolution.


Squares: Blasius solution; Circles: Spalart & Strelets (2000) DNS; diamonds:
under-resolved DNS; dashed line: -model; dash-dotted line: dynamic Smagorinsky; full line: truncated Navier-Stokes.
92

25
20

u+

15
10
5
0

10

10

10

10

y+

Figure 6.6: Turbulent boundary layer mean streamwise velocity in wall units at
x = 7 for LES-V at 1% of DNS resolution. Squares: log law of the wall; circles:
u+ = y + ; diamonds: under-resolved DNS; dashed line: -model; dash-dotted line:
dynamic Smagorinsky; full line: truncated Navier-Stokes.

93

0.2

0.18

0.18

0.16

0.16

0.14

0.14

0.12

0.12

0.1

0.1

0.2

0.08

0.08

0.06

0.06

0.04

0.04

0.02

0.02

0.5

1.5

2.5

0.2

0.2

0.18

0.18

0.16

0.16

0.14

0.14

0.12

0.12

0.1

0.1

0.08

0.08

0.06

0.06

0.04

0.04

0.02

0.02

0.5

1.5

(c) -model.

0.5

1.5

2.5

(b) dynamic Smagorinsky.

(a) under-resolved DNS.

2.5

0.5

1.5

2.5

(d) truncated Navier-Stokes.

Figure 6.7: Proles in LES-V at 1% of DNS resolution (lines), and in DNS (symbols). Line, circles: U; dashed line, diamonds: u /umax . From left to right, plots
spaced by 1.5: x = 2.5, x = 3.5.

94

0.25

0.2

0.15

0.1

0.05

4
x

(a) umax (x).


0.25

0.2

0.15

0.1

0.05

4
x

(x).
(b) vmax

Figure 6.8: Maximum RMS uctuation velocity in LES-V at 1% of DNS resolution.


Circles: Spalart & Strelets (2000) DNS; diamonds: under-resolved DNS; dashed
line: -model; dash-dotted line: dynamic Smagorinsky; full line: truncated NavierStokes.
95

Chapter 7
Conclusions
7.1

Summary and Conclusions

Large eddy simulations of a laminar separation bubble ow over a at plate


with a Reynolds number (105 ) and pressure distribution comparable to those seen
on airfoils of small unmanned aircraft and on blades of low pressure turbines
were performed using three high-order codes: a sixth-order compact nite difference compressible large eddy simulation solver provided by the NASA Stanford
Center for Turbulence Research (CTR code) and two Fourier-Chebyshev pseudospectral incompressible Navier-Stokes solvers. The performance of each simulation
was evaluated against benchmark DNS data focusing on two quantities of critical importance to airfoil and blade designers: time-averaged pressure (Cp ) and
skin friction (Cf ) predictions used in lift and drag calculations. Boundary layer
thicknesses, mean velocity proles, and RMS uctuation velocity proles were also
investigated.
Agreement obtained with two spectral large eddy simulation tools with explicit
subgrid-scale models demonstrates that accurate large eddy simulation of laminar
separation bubble ows are attainable with as low as 1% of DNS resolution. The
level of agreement obtained with explicitly-modeled LES is far superior to RANS
results to date and has not been previously demonstrated for laminar separation
bubble ows at such low resolution. Strong departures from the DNS benchmark are observed in the no-model spectral simulations, indicating that explicit
96

subgrid-scale models are required in low-dissipation low-resolution cases. In contrast, no-model predictions given by the CTR code were in better agreement with
the CTR DNS due to its use of an explicit stability-enhancing lter providing
enough dissipation without the use of a subgrid-scale model.
In fact, the addition of an explicit subgrid-scale model in the CTR code deteriorated agreement with the CTR DNS benchmark. Numerical dissipation due
to ltering was rst estimated by comparing energy decay rates in the turbulent
boundary layer region downstream of reattachment to new simulations with no
ltering active, but higher physical viscosity. The equivalent numerical viscosity
was estimated to be of the same order of magnitude as physical viscosity and comparable to the average eddy viscosity provided by the dynamic Smagorinsky model
in the same region. Numerical dissipation due to ltering was then conrmed to
play an important role equivalent to an implicit subgrid-scale model using a novel
method to quantify it based solely on post-processing available consecutive velocity
elds.
To assess the performance of dierent subgrid-scale models without the eects
of numerical dissipation seen in the CTR code results, a spectral solver with negligible numerical dissipation was used to perform large eddy simulation at two resolutions corresponding to 3% and 1% of the benchmark DNS by Spalart & Strelets
(2000). Three explicit subgrid-scale models were compared to a no-model baseline
case or under-resolved direct numerical simulation (UDNS) for these two resolutions. Simulations without an explicit subgrid-scale model (UDNS) failed to
capture accurately the benchmark Cf and Cp at 3% resolution, and even failed
to capture qualitative features of the expected ow on the coarsest mesh where
its results were contaminated by excessive energy accumulating in the smallest
resolved scales. The addition of explicit subgrid-scale models markedly improved
97

agreement with the benchmark DNS. At 3% of DNS resolution, all explicitlymodeled LES tested recover all important features of the ow almost exactly, with
only minor discrepancies. The truncated Navier-Stokes approach performed the
best on the coarsest grid, recovering the overall skin friction, pressure, boundary
layer thicknesses, and velocity proles of the DNS benchmark most accurately. The
model performed the second best, but like the dynamic Smagorinsky model, it
did not capture the onset of the peak negative skin friction or the reattachment
point as accurately at 1% of DNS resolution. However, both recovered the turbulent skin friction downstream of reattachment almost exactly. All results obtained
with explicit subgrid-scale models are self-consistent as well as consistent amongst
each other for the two resolutions and codes tested. The observed grid independence strengthens the validity of these subgrid-scale modeling approaches. Results
obtained using a dierent spectral solver corroborate the above-mentioned conclusions with only minor discrepancies attributed to dierences in implementation
and boundary conditions.
That the physics of the problem were well captured at this low resolution
is encouraging for applications with body-tted meshes and higher operating
Reynolds numbers. However, dissipative schemes may deteriorate large eddy simulation predictions as illustrated in the CTR large eddy simulation results, meaning
that higher order methods like nite element or discontinuous Galerkin may be
required. Large eddy simulation coupled with such high-order methods should be
able to converge to DNS results at resolutions two orders of magnitude lower than
would otherwise be required for DNS, providing signicant computational savings
that could soon enable design optimization for applications that are dominated
by the unsteady eects of separation and transition. For applications where such
high order methods are not available or are not practical, eorts should be made to
98

remove sources of numerical dissipation where possible. For example, the highest
order central dierence ux approximations available should be chosen instead
of other more dissipative schemes like upwinding or essentially non-oscillatory
schemes. Once a statistically-steady state is reached on a coarse grid, numerical
dissipation should be quantied in dierent subvolumes using the method developed by Schranner et al. (2015), especially in transitional and turbulent regions of
the ow. If necessary, the resolution should be increased locally in these subvolumes in order to decrease the eective numerical viscosity until it is an order of
magnitude smaller than the subgrid-scale viscosity of an established subgrid-scale
model. This approach should help the user converge to a solution more quickly. It
is also less costly than typical grid-renement studies because long averaging times
are not necessary. However, it does not provide the same invaluable convergence
rate information.

7.2

Future Work

The consistently good agreement obtained with the truncated Navier-Stokes


at 1% of the DNS resolution may be evidence that this approach more accurately
represents the underlying physics than the traditional approach to large eddy simulation using the ltered Navier-Stokes equations with a modeled subgrid-scale
stress tensor. The truncated Navier-Stokes approach thus warrants further study.
Namely, it should be extended to and tested on other more exible high order
codes for both canonical ows as well as more complex geometries where DNS are
available to assess its accuracy. To better understand its underlying dynamics, the
evolution of the energy spectra and transfer terms should be plotted and compared
to a no-model simulation as well as a DNS of the same ow if available.

99

The simple method to quantify numerical dissipation proposed by


Schranner et al. (2015) should be investigated as a tool to gauge whether explicit
subgrid-scale models will yield worthwhile improvements in results for a given grid
and a nominally under-resolved solution with no model for commercial package
users. Moreover, this method could be used on-the-y to help determine when and
where subgrid-scale models should be active, or when ltering is necessary in the
case of the truncated Navier-Stokes approach. Another application could be found
in solution-adaptive mesh renement in which areas where numerical dissipation
is computed and found to be higher than an acceptable threshold are rened.
The capability of large eddy simulation to succesfully predict time-averaged
quantities at resolutions on the order of 1% of DNS should be tested on blades
and airfoils at or near their operating Reynolds number in congurations for which
DNS or experimental data is available. The same simulation setups should also
be used to compare low resolution large eddy simulation results with dierent wall
models active, and with hybrid RANS-LES formulations so their impact on the
quality of results can be assessed.

100

Reference List
Alam M, Sandham N (2000) Direct numerical simulation of short laminar separation bubbles with turbulent reattachment. Journal of Fluid Mechanics 410:128.
Alving AE, Fernholz HH (1996) Turbulence measurements around a mild separation bubble and downstream of reattachment. Journal of Fluid Mechanics 332:297328.
Anderson BW, Domaradzki JA (2012) A subgrid-scale model for large-eddy simulation based on the physics of interscale energy transfer in turbulence. Physics
of Fluids 24:065104.
Basara B, Krajnovic S, Girimaji S, Pavlovic Z (2011) Near-wall formulation of the
partially averaged navier stokes turbulence model. AIAA Journal 49:26272636.
Bogey C, Bailly C (2006) Large eddy simulations of transitional round jets:
Inuence of the Reynolds number on ow development and energy dissipation.
Physics of Fluids 18:065101.
Bose ST, Moin P (2014) A dynamic slip boundary condition for wall-modeled
large-eddy simulation. Physics of Fluids (1994-present) 26:.
Burgmann S, Bcker C, Schrder W (2006) Scanning PIV measurements of a
laminar separation bubble. Experiments in Fluids 41:319326.
Burgmann S, Dannemann J, Schrder W (2007) Time-resolved and volumetric
PIV measurements of a transitional separation bubble on an SD7003 airfoil.
Experiments in Fluids 44:609622.
Burgmann S, Schrder W (2008) Investigation of the vortex induced unsteadiness of a separation bubble via time-resolved and scanning PIV measurements.
Experiments in Fluids 45:675691.
Cadieux F, Domaradzki JA, Sayadi T, Bose T (2014) DNS and LES of laminar
separation bubbles at moderate Reynolds numbers. Journal of Fluids Engineering 136.
101

Cadieux F, Domaradzki JA, Sayadi T, Bose T, Duchaine F (2012) DNS and LES
of separated ows at moderate Reynolds numbers In Proceedings of the 2012
Summer Program, pp. 7786. Center for Turbulence Research.
Canuto C, Hussaini MY, Quarteroni A, Zang TA (2007) Spectral Methods : Fundamentals in Single Domains Springer, Dordrecht.
Castiglioni G (2015) Numerical Modeling of Separated Flows at Moderate
Reynolds Numbers Appropriate for Turbine Blades and Unmanned Aero Vehicles Ph.D. diss., University of Southern California.
Castiglioni G, Domaradzki J, Pasquariello V, Hickel S, Grilli M (2014) Numerical
simulations of separated ows at moderate reynolds numbers appropriate for
turbine blades and unmanned aero vehicles. International Journal of Heat and
Fluid Flow 49:91 99 8th Symposium on Turbulence & Shear Flow Phenomena
(TSFP8).
Deck S (2012) Recent improvements in the zonal detached eddy simulation (zdes)
formulation. Theoretical and Computational Fluid Dynamics 26:523550.
Deck S, lie Weiss P, Pamis M, Garnier E (2011) Zonal detached eddy simulation of a spatially developing at plate turbulent boundary layer. Computers
& Fluids 48:1 15.
Diamessis P, Lin Y, Domaradzki JA (2008) Eective numerical viscosity in spectral
multidomain penalty method-based simulations of localized turbulence. Journal
of Computational Physics 227:81458164.
Domaradzki JA, Metcalfe RW (1987) Stabilization of laminar boundary layers by
compliant membranes. Physics of Fluids 30:695705.
Domaradzki JA, Radhakrishnan S (2005) Eective eddy viscosities in implicit
modeling of decaying high Reynolds number turbulence with and without rotation. Fluid Dynamics Research 36:385406.
Domaradzki JA, Xiao Z, Smolarkiewicz PK (2003) Eective eddy viscosities in implicit large eddy simulations of turbulent ows. Physics of Fluids 15:38903893.
Domaradzki JA, Loh KC, Yee PP (2002) Large eddy simulations using the subgridscale estimation model and truncated navier-stokes dynamics. Theoretical and
Computational Fluid Dynamics 15:421450.
Dubois T, Domaradzki J, Honein A (2002) The subgrid-scale estimation model
applied to large eddy simulations of compressible turbulence. Physics of Fluids 14:17811801.
102

Egorov Y, Menter F, Lechner R, Cokljat D (2010) The scale-adaptive simulation


method for unsteady turbulent ow predictions. part 2: Application to complex
ows. Flow, Turbulence and Combustion 85:139165.
Eisenbach S, Friedrich R (2008) Large-eddy simulation of ow separation on an
airfoil at a high angle of attack and Re = 105 using Cartesian grids. Theor.
Comp. Fluid Dyn. 22:213225 10.1007/s00162-007-0072-z.
Erlebacher G, Hussaini M, Speziale C, Zang T (1992) Toward the large-eddy simulation of compressible turbulent ows. Journal of Fluid Mechanics 238:155185.
Fletcher CAJ (1991) Computational techniques for fluid dynamics, Vol. 1 Springer
Verlag.
Gaster M (1963) On the stability of parallel ows and the behaviour of separation
bubbles Ph.D. diss., University of London.
Girimaji S, Abdol-Hamid K (2005) chapter Partially-Averaged Navier Stokes
Model for Turbulence: Implementation and Validation Aerospace Sciences Meetings. American Institute of Aeronautics and Astronautics 0.
Guermond J, Minev P, Shen J (2006) An overview of projection methods for
incompressible ows. Computer Methods in Applied Mechanics and Engineering 195:6011 6045.
Hadi I, Hanjali K (2000) Separation-induced transition to turbulence: Secondmoment closure modelling. Flow, Turbulence and Combustion 63:153173.
Hggmark C (2000) Investigation of disturbances developing in a laminar separation bubble ow Ph.D. diss., Royal Institute of Technology, Sweden.
Hain R, Khler CJ, Radespiel R (2009) Dynamics of laminar separation bubbles
at low-Reynolds-number aerofoils. Journal of Fluid Mechanics 630:129153.
Herbst A, Henningson D (2006) The inuence of periodic excitation on a turbulent
separation bubble. Flow, Turbulence and Combustion 76:121.
Horton H (1968) Laminar separation bubbles in two and three dimensional incompressible ow Ph.D. diss., University of London.
Howard R, Alam M, Sandham N (2000) Two-equation turbulence modelling of a
transitional separation bubble. Flow, Turbulence and Combustion 63:175191.
Hu H, Yang Z, Igarashi H (2007) Aerodynamic hysteresis of a low-Reynolds-number
airfoil. Journal of Aircraft 44:20832085.

103

Jones LE, Sandberg RD, Sandham ND (2008) Direct numerical simulations of


forced and unforced separation bubbles on an airfoil at incidence. Journal of
Fluid Mechanics 602:175207.
Jones L, Sandberg R, Sandham N (2010) Stability and receptivity characteristics of a laminar separation bubble on an aerofoil. Journal of Fluid Mechanics 648:257296.
Kojima R, Nonomura T, Oyama A, Fujii K (2013) Large-eddy simulation of
Low-Reynolds-Number ow over thick and thin NACA airfoils. Journal of Aircraft 50:187196.
Kravchenko A, Moin P (1997) On the eect of numerical errors in large eddy
simulations of turbulent ows. Journal of Computational Physics 131:310 322.
Lele SK (1992) Compact nite dierence schemes with spectral like resolution.
Journal of Computational Physics 103:1642.
Lin JCM, L.Pauley L (1996) Low-Reynolds-number separation on an airfoil. AIAA
Journal 34:15701577.
Mani A (2012) Analysis and optimization of numerical sponge layers as a nonreective boundary treatment. Journal of Computational Physics 231:704 716.
Marxen O, Henningson D (2011) The eect of small-amplitude convective disturbances on the size and bursting of a laminar separation bubble. Journal of Fluid
Mechanics 671:133.
Marxen O, Lang M, Wagner S (2003) A combined experimental/numerical study
of unsteady phenomena in a laminar separation bubble. Flow, Turbulence and
Combustion 71:133146.
Marxen O, Rist U (2004) Direct and large eddy simulation of the transition process
in a laminar separation bubble In Proceedings of the 5th Workshop on Direct
and Large-Eddy Simulation, pp. 231240, Munich, Germany. Kluwer Academic
Publishers.
Marxen O, Rist U (2010) Mean ow deformation in a laminar separation bubble:
separation and stability characteristics. Journal of Fluid Mechanics 660:3754.
McMullan W, Page G (2012) Towards large eddy simulation of gas turbine compressors. Progress in Aerospace Sciences 52:30 47 Applied Computational
Aerodynamics and High Performance Computing in the {UK}.

104

Menter F, Egorov Y (2010) The scale-adaptive simulation method for unsteady


turbulent ow predictions. part 1: Theory and model description. Flow, Turbulence and Combustion 85:113138.
Na Y, Moin P (1998) Direct numerical simulation of a separated turbulent boundary layer. Journal of Fluid Mechanics 374:379405.
Nagarajan S (2004) Leading edge eects in bypass transition Ph.D. diss., Stanford
University.
Nagarajan S, Lele SK, Ferziger JH (2003) A robust high-order compact method
for large eddy simulation. Journal of Computational Physics 191:392419.
Nagarajan S, Lele SK, Ferziger JH (2007) Leading-edge eects in bypass transition.
Journal of Fluid Mechanics 572:471504.
Nicoud F, Toda HB, Cabrit O, Bose S, Lee J (2011) Using singular values to
build a subgrid-scale model for large eddy simulations. Physics of Fluids (1994present) 23:.
Orszag SA, Kells LC (1980) Transition to turbulence in plane poiseuille and plane
couette ow. Journal of Fluid Mechanics 96:159205.
Papanicolaou EL, Rodi W (1999) Computation of separated-ow transition using
a two-layer model of turbulence. Journal of Turbomachinery 121:7887.
Piomelli U (1999) Large-eddy simulation: achievements and challenges. Progress
in Aerospace Sciences 35:335 362.
Piomelli U (2008) Wall-layer models for large-eddy simulations. Progress in
Aerospace Sciences 44:437 446 Large Eddy Simulation - Current Capabilities
and Areas of Needed Research.
Piomelli U, Balaras E (2002) Wall-layer models for large-eddy simulations. Annual
Review of Fluid Mechanics 34:349374.
Roberts S, Yaras M (2005) Large-eddy simulation of transition in a separation
bubble. Journal of Fluids Engineering 128:232238.
Sagaut P (2006) Large Eddy Simulation for Incompressible Flows Springer, Berlin,
Germany.
Sagaut P, Deck S (2009) Large eddy simulation for aerodynamics: status and perspectives. Philosophical Transactions of the Royal Society of London A: Mathematical, Physical and Engineering Sciences 367:28492860.

105

Sayadi T, Moin P (2012) Large eddy simulation of controlled transition to turbulence. Physics of Fluids 24:114103.
Schranner FS, Domaradzki JA, Hickel S, Adams NA (2015) Assessing the numerical dissipation rate and viscosity in numerical simulations of uid ows. Computers & Fluids 114:84 97.
Shur ML, Spalart PR, Strelets MK, Travin AK (2008) A hybrid RANS-LES
approach with delayed-DES and wall-modelled {LES} capabilities. International
Journal of Heat and Fluid Flow 29:1638 1649.
Skote M, Henningson D (2002) Direct numerical simulation of a separated turbulent boundary layer. Journal of Fluid Mechanics 471:107136.
Skote M, Henningson D, Henkes R (1998) Direct numerical simulation of selfsimilar turbulent boundary layers in adverse pressure gradients. Flow, Turbulence and Combustion 60:4785.
Sohn KH, Shyne RJ, DeWitt KJ (1998) Experimental investigation of boundary
layer behavior in a simulated low pressure turbine National Aeronautics and
Space Administration, Lewis Research Center.
Spalart PR (2009) Detached-eddy simulation. Annual Review of Fluid Mechanics 41:181202.
Spalart P (2006) Topics in detached-eddy simulation In Groth C, Zingg D, editors,
Computational Fluid Dynamics 2004, pp. 312. Springer Berlin Heidelberg.
Spalart P, Strelets M (2000) Mechanisms of transition and heat transfer in a
separation bubble. Journal of Fluid Mechanics 403:329349.
Spalart P, Watmu J (1993) Experimental and numerical study of a turbulent
boundary layer with pressure gradients. Journal of Fluid Mechanics 249:337371.
Spedding G, McArthur J (2010) Span eciencies of wings at low Reynolds numbers. Journal of Aircraft 47:120128.
Stolz S, Adams NA, Kleiser L (2001) An approximate deconvolution model for
large-eddy simulation with application to incompressible wall-bounded ows.
Physics of Fluids 13:9971015.
Tantikul T, Domaradzki J (2010) Large eddy simulations using truncated NavierStokes equations with the automatic ltering criterion. Journal of Turbulence 11:124.

106

Tantikul T, Domaradzki J (2011) Large eddy simulations using truncated NavierStokes equations with the automatic ltering criterion: Reynolds stress and
energy budgets. Journal of Turbulence 12:125.
Wilcox DC (2006) Turbulent Modeling for CFD DCW Industries, La Canada,
California.
Wilson P, Pauley L (1998) Two-and three-dimensional large-eddy simulations of a
transitional separation bubble. Physics of Fluids 10:29322940.
Wu X, Moin P (2010) Transitional and turbulent boundary layer with heat transfer.
Physics of Fluids 22:85105.
Xu T, Sullivan P, Paraschivoiu M (2010) Fast large-eddy simulation of low
Reynolds number ows over a NACA0025. Journal of Aircraft 47:328333.
Yang Z, Voke P (2001) Large-eddy simulation of boundary-layer separation
and transition at a change of surface curvature. Journal of Fluid Mechanics 439:305333.
Yarusevych S, Sullivan PE, Kawall JG (2006) Coherent structures in an airfoil
boundary layer and wake at low Reynolds numbers. Physics of Fluids 18:044101.

107

You might also like