You are on page 1of 21

AIAA 2010-4600

40th Fluid Dynamics Conference and Exhibit


28 June - 1 July 2010, Chicago, Illinois

40th Fluid Dynamics Conference and Exhibit, 28th June 1st July, 2010, Chicago, Illinois

Numerical investigation of the effect of free-stream


turbulence on laminar boundary-layer separation
Wolfgang Balzer, and Hermann F. Fasel
Department of Aerospace and Mechanical Engineering,
The University of Arizona, Tucson, Arizona 85721
Direct numerical simulations (DNS) are employed to investigate the effect of free-stream
turbulence (FST) on laminar boundary-layer separation and separation control. To model
the effects of FST, a numerical method first proposed by Jacobs1 for the generation of
isotropic grid turbulence is used. For a laminar separation bubble on a flat plate, which
was investigated in wind-tunnel experiments by Gaster,2 it is shown that using low levels
of FST in numerical simulations can improve the agreement between experiments and
DNS. The characteristics of the fluctuations inside the boundary layer prior to separation
agree very well with those reported in experiments and other numerical simulations. Even
for the highest turbulent intensity in the free stream (2.5%), which we investigated, the
boundary layer is found to separate from the surface and undergo transition to turbulence
in the separated shear layer above the wall. In addition, time-dependent flow analysis
and comparison between DNS results and linear stability theory reveal that the linear
shear-layer instability mechanism (Kelvin-Helmholtz instability) is not bypassed but can
be detected in all cases. DNS of active flow control using harmonic blowing and suction
through a narrow spanwise slot is shown to significantly reduce separation and delay flow
transition. For increased levels of FST, however, the high-amplitude, two-dimensional
disturbance waves introduced by the forcing are seen to interact with three-dimensional
disturbances inside the boundary layer which results in an accelerated transition process
when compared to the uncontrolled case.

I.

Introduction

The free-stream turbulent intensity can be relatively low for wind turbine or free-flight applications
(< 0.5%), but also very high as in turbomachinery flow. For the latter, turbulent intensities as high as 20%
have been reported.3 Historically, free-stream turbulence (FST) has made it very difficult to identify linear
stages of the transition process. Experimental evidence of Tollmien-Schlichting (TS) waves, for example,
was not found until Schubauer and Skramstad4 managed to reduce the level of turbulent fluctuations in the
free stream to unusually low values ( 1%). The success of Schubauers experiments could in part be
attributed to suggestions by Dryden5 who carried out experiments for a much more turbulent free stream.
He reported slow irregular u-velocity fluctuations of large amplitude in the laminar region of the boundary
layer that were induced by the turbulence in the free stream. In fact, transition in wind-tunnel experiments
typically seems to be preceded by such a low-frequency streaky modulation of the boundary layer. Today,
these streaks are commonly referred to as the Klebanoff modes after P. S. Klebanoff who investigated
this phenomenon6, 7 and described them as a periodic thickening/thinning of the boundary layer. From
other experimental studies investigating the effect of FST it has been established that in addition to its low
frequency and high amplitude, urms = O(0.1U ), the Klebanoff mode is also characterized
by a distinct

spanwise scaling of O(299 499 ) and an algebraic streamwise growth proportional to x.8, 9, 10
Naturally, the effect of FST is equally important in the presence of streamwise pressure gradients (see
Mayle,11 for example). Adverse pressure gradients can lead to the formation of a laminar separation bubble
Research

Assistant.
AIAA member.

Professor,

1 of 21
Institute
of Aeronautics
and Astronautics
Paper 2010-4600
Copyright 2010 by the authors. Published by the American
American Institute
of Aeronautics
and Astronautics,
Inc., with permission.

(LSB). In this case, the flow detaches from the solid surface, undergoes transition, and eventually reattaches to the surface as a transitional or turbulent flow. LSBs are a common flow phenomenon observed
in turbomachinery, high-lift multi-element airfoil configurations and diffusers. Elevated levels of FST have
been reported to reduce the length of the separation bubble12, 13 or even prevent separation completely by
transitioning the flow upstream of the separation location.14 However, early transition is not necessarily
beneficial. The significant increase in skin-friction and heat-transfer rate associated with turbulent flow can
have a detrimental effect on efficiency and hardware durability.15
The existence of low-frequency, large-amplitude streaks in the boundary layer and in the separated shear
layer has been confirmed in the experiments of Haggmark16 and Volino.13 In the experiments by Haggmark,
grid-generated turbulence of 1.5% intensity was considered for a separation bubble developing on a flat
plate. Using smoke visualization techniques, he found that there is no strong evidence of two-dimensional
waves which typically can be observed for separation bubbles developing in an undisturbed environment. In
addition, he performed a spectral analysis and observed that lower frequencies are dominant in the case of
increased FST. Volinos experimental investigations focused on the effect of low and high FST levels on the
suction side separation of a low-pressure turbine blade. For a low Reynolds number Re = 25, 000 (based on
the blades suction surface length and exit velocity), he reported that, even in the case of high levels of FST
(9%), breakdown to turbulence occured in the separated shear layer above the wall and that the flow did
not reattach upstream of the trailing edge.
For low FST a generally accepted route of transition for a flat-plate boundary layer consists of four
discrete processes: receptivity, linear growth of 2-D TS waves, secondary instabilities, and breakdown to
turbulence. Since the presence of a highly disturbed free stream introduces finite non-linear disturbances in
the laminar boundary layer, the first stages of the transition process (linear growth) may be bypassed. The
terminology of bypass transition was first introduced by Morkovin.17 Lately, the term continuous mode
transition is also used18 arguing that bypass transition is preceded by the amplification of continuous modes
of the Orr-Sommerfeld spectrum rather than discrete modes (TS waves). Transition initiated by continuous modes can be seen in connection with transient growth which is associated with the non-normality
of the linearized Navier-Stokes operator. Although individual continuous modes decay exponentially, they
are non-orthogonal and their superposition can lead to significant transient growth of the streamwise velocity component. Optimal disturbance growth (i.e., the optimal superposition) is observed for a spanwise
periodic array of steady, counter-rotating streamwise vortices with a distinct spanwise scaling.19, 20 If the
initial disturbance amplitudes are sufficiently large and a critical energy threshold is exceeded, transition is
initiated via nonlinear interactions. Wu et al.21, 22 employed 3-D DNS to investigate the separated flow in
a low-pressure turbine stage under the influence of turbulent, periodically passing wakes. They studied the
connection between classical linear growth mechanisms (Tollmien-Schlichting, Kelvin-Helmholtz) and transient growth, concluding that transient growth plays an important role in this context. For a zero pressure
gradient, Reshotko23 showed that transient growth of steady spanwise disturbances could help to explain
bypass transition. The potential of delaying separation via a secondary instability of (algebraically) growing
streamwise vortices, e.g., introduced by steady vortex generator jets or surface roughness, has motivated
theoretical investigations in the scope of transient growth.24 Due to their distinct spanwise modulation and
typical wall-normal u-velocity distributions, Klebanoff modes are often misconceived as streamwise vortices
(also introduced as such by Klebanoff 6 ). Contrary, the disturbances rather take the form of elongated streaks
in the laminar boundary layer with comparatively low levels of streamwise vorticity.18, 25 Consequently, from
a transient growth point-of-view, the K-modes are not optimal disturbances. In addition, the question of
how the spanwise scaling for the free-stream induced K-modes compares to the optimal spacing found from
transient growth theory is unanswered. In both cases, however, this spanwise scaling seems to be an intrinsic
property of the boundary layer.
For our investigations we employ highly-accurate direct numerical simulations (DNS) to study the effect
of free-stream turbulence on laminar boundary-layer separation. In section II, we introduce the governing
equations and numerical method. Validation of a numerical model for the generation of realistic freestream disturbances is presented in section III, followed by the description of the simulation setup in section
IV. Results for the uncontrolled separation bubble with and without free-stream turbulence as well as an
investigation of flow control are presented in section V. Finally, conclusions are provided in section VI.

2 of 21
American Institute of Aeronautics and Astronautics Paper 2010-4600

II.

Governing Equations and Numerical Method

The governing equations are the three-dimensional, incompressible, time-dependent Navier-Stokes (NS)
equations. In the present work, the Navier-Stokes equations are solved in a vorticity-velocity formulation
which can be obtained by taking the curl of the momentum equations and thus eliminating the pressure
term,

1 2
+ (v ) ( ) v =
,
(1)
t
Re
where both the vorticity vector, defined as = v = [x , y , z ]T , and the velocity vector, v = [u, v, w]T ,
are solenoidal. The velocity field, v, is determined by solving the vector Poisson equation
2 v = .

(2)

The global Reynolds number in Eq. 1 is defined as Re = L U /, where L is a reference length, U is


the free-stream velocity, and is the kinematic viscosity.
Rist and Fasel26 point out that the employed formulation is in the so-called conservative form. It
can be shown that this is the case if in fact the vorticity field is solenoidal. Divergence-free velocity- and
vorticity fields and therefore the conservation of either quantity are, however, not necessarily guaranteed by
the numerical solution, in particular since both the continuity equation and the definition of vorticity do not
explicitly appear in the set of governing equations, Eqs. 1 and 2. (Note, however, that v = 0 has been used
to derive these equations). In other words, vorticity could be dispersed or accumulated. Using the maximum
principle for Laplaces equation, Fasel27 showed that the continuity equation can still be satisfied inside the
integration domain if it is satisfied on the boundaries of the domain. For any numerical solution this would
result in an error which is maximal on the boundaries. These difficulties have led to restrictions in the use
of the vorticity-velocity formulation in the form presented above. Marxen,28 for example points out that no
disturbances are possible at the inflow boundary. Allowing for unsteady, non-uniform inflow disturbances
was, however, a fundamental requirement for the numerical results presented in the current work. Several
test simulations and validation cases revealed that it would be necessary to use a reformulation of the
governing equations in cases where time-dependent inflow conditions are considered. Although the velocityand vorticity fields enforced at the inflow boundary can always be chosen such that they satisfy continuity to
a high degree, it is important that the numerical scheme guarantees continuity inside the integration domain.
The solution methodology proposed by Davies and Carpenter29 is based on solving = 0 explicitly. We
adopted this approach by replacing the spanwise vorticity transport equation (Eq. 1ez ) with the constraint
that the vorticity field must be solenoidal at all times. In other words, the spanwise vorticity z is obtained
by solving
z

=

.
(3)
z
x
y
The flow field is assumed to be periodic in the spanwise z-direction. Employing a pseudo-spectral approach, each variable is expanded in a Fourier series and represented by a total of 2K +1 Fourier modes. The
Fourier representation of the spanwise direction reduces the governing equations to a set of two-dimensional
equations for each Fourier mode. The set of 2-D problems is solved in parallel using 2K + 1 processors on a
modern supercomputing platform. For the calculation of the nonlinear terms, the flow field is transformed
from spectral to physical space (and back) before each Runge-Kutta sub step which requires redistributing of
the entire three-dimensional arrays among the processors. This extensive inter-processor communication is
realized using the message passing interface (MPI). To avoid aliasing errors, the nonlinear terms in physical
space are computed on 3K spanwise collocation points.
A.

Generation of inflow disturbances

Free-stream turbulence is often modelled by introducing random velocity disturbances at the inflow boundary.
However, in the method proposed by Jacobs1 these velocities are not completely random but are designed
such as to satisfy continuity, represent isotropic turbulence conditions and satisfy a specified energy spectrum.
We adapted this method and implemented it in our current Navier-Stokes solver. The method is based on
an expansion of the disturbance velocity, v , in terms of its Fourier coefficients,
v
(k, t) eikx ,

v (x, t) =
k

3 of 21
American Institute of Aeronautics and Astronautics Paper 2010-4600

(4)

where k is the wave-vector with length k =

kx2 + ky2 + kz2 . The inflow boundary condition is then a super-

position of a steady, undisturbed base flow VB (typically a Blasius solution) and the unsteady disturbance
velocity:
V(x0 , y, z, t) = VB (y) + v (x0 , y, z, t) .
(5)
In the same manner, the inflow boundary condition for the vorticity field is
(x0 , y, z, t) = B (y) + (x0 , y, z, t) ,

(6)

where is determined from the disturbance velocity field as = v .


Turbulent fluctuations are only introduced at the inflow boundary plane (x = x0 , y, z) and hence kx x is
replaced by t assuming that the disturbances are convected with the speed of the free stream (Taylors
hypothesis). Note that we use the less common symbol to denote the circular frequency since shall be
exclusively used to refer to vorticity and its components.
v (x0 , y, z, t) =

v
(ky , kz , ) eiky y eikz z eit .

kz

(7)

ky

The implementation of the spanwise Fourier cosine and sine modes, eikz z , is straightforward since the numerical model assumes periodicity of the flow field in z-direction. However, the expansion in wall-normal
direction requires a modification due to the presence of the wall. Jacobs1 suggests to use local solutions in
the continuous spectrum of the linearized Navier-Stokes equations as an alternative basis. The linear NS
equations are the Orr-Sommerfeld (OS) and Squire (SQ) equation:
d2
2
dy 2

d2
2
dy 2

iRe (U c)

d2
2
dy 2

d2 U
dy 2

v = 0 ,

iRe (U c) = iRe

(8a)
dU
v ,
dy

(8b)

where U = U (y) is the streamwise velocity of a parallel baseflow, and c = / is the phase speed of
a travelling disturbance for which v = v(y) and = (y) are the wall-normal velocity and wall-normal
vorticity amplitude distribution. The travelling waves are of the form
v = v(y)ei(x+zt) ,

= (y)ei(x+zt) ,

(9)

and = 2 + 2 is the resultant wavenumber. The general solution of Eqs. 8 gives rise to a discrete spectrum (e.g. Tollmien-Schlichting modes) and a continuous spectrum. For wall-bounded flows the boundary
conditions of the eigenvalue problem are typically chosen as v = d
v /dy = = 0 at solid walls and in the
free stream. In order to obtain oscillatory free-stream modes, the free-stream boundary conditions can be
relaxed by assuming that v, d
v /dy and are bounded, but not zero, as y . Assuming solutions of the
form eiky y the eigenvalue problem, Eq. 8a, reduces to a fourth-order boundary value problem with constant
coefficients for which the eigenvalues can be determined explicitly (see Jacobs1 or Grosch and Salwen30 for
details),
i
=
Re2 + 4 ky2 iRe Re .
(10)
2
In order to remove unwanted resonances between the Squire equation and the Orr-Sommerfeld equation,
Eqs. 8a and 8b are usually decoupled by neglecting the influence of the right-hand-side forcing term. The
continuity equation in the form iu + v/y + iw = 0 and the definition of the wall-normal vorticity
= iw iu are then used to find expressions for the streamwise and spanwise velocity components
u=

i
2

v
+
y

w=

i
2

v

y

(11)

A typical mode from the continuous spectrum of the simplified Orr-Sommerfeld equation is plotted in
Fig. 1. The solution shows good agreement with that obtained by Jacobs.1 From Fig. 1, it can be seen that
the continuous modes obtained using this approach are sinusoidal in the free stream, but naturally decay to
4 of 21
American Institute of Aeronautics and Astronautics Paper 2010-4600

vr

0.5

-0.5

-1
0

10

12

14

y
Figure 1. Comparison of numerical solutions for the real part, vr , of the spatial eigenfunction for Re = 639,
ky = , = 0.5 and the base flow Ub (). Results are presented for various spatial resolutions with Ny = 129
points ( ), 193 points ( ), 257 points ( ), 512 points () in comparison to results by Jacobs1
( )

low amplitudes inside the boundary layer. This behavior makes them ideal candidates for the generation of
inflow disturbances. Note that solutions to the simplified Squire equation look very similar to the OS modes
(only differ by a phase shift).
Using the continuous modes as a basis for the expansion in wall-normal direction, the disturbance velocity
components at the inflow can be written as
u(x0 , y, z, t) =

kz

ky

i
v
i
+ B 2 eikz eit ,
2
y

A
v eikz eit ,

v(x0 , y, z, t) =

kz

(12b)

ky

w(x0 , y, z, t) =

(12a)

kz

ky

v
i
i
B 2 eikz eit .
2 y

(12c)

The complex amplitudes A = A(, ky , kz ) and B = B(, ky , kz ) are defined as:


A = f ei1 cos ,

B = f ei2 sin ,

(13)

where 1 , 2 and are uniformly distributed random numbers which provide random distribution and
random phase among the the Fourier coefficients. The additional scaling, f , can be linked to the turbulent
kinetic energy by considering
q2
1
=
2
2

u
i u
i =
k

1
2

(AA + BB ) =
k

1
2

f2 .

(14)

The turbulent kinetic energy can also be written in terms of the three-dimensional energy spectrum.
q2
=
2

(15)

Equations 14 and 15 lead to a definition for f =


Karman energy spectrum in the form,
E(k) =

E(|k|)k

E(k)dk
0

2E(k)k. Following Jacobs suggestion we use the von

1.196(kL11 )4
17

[0.558 + (kL11 )2 ] 6

L11 T u2 ,

5 of 21
American Institute of Aeronautics and Astronautics Paper 2010-4600

(16)

where k was normalized by the free-stream integral length scale L11 and T u is the free-stream turbulent
intensity defined as
Tu =

1
3

(u2 ) + (v 2 ) + (w2 )
U

(17)

Figure 2 shows a typical energy spectrum. It can be deduced that the choice of L11 controls the size
(wavenumber) of the most energetic structures and that T u controls the amplitude levels of these disturbances. The energy spectrum can be envisioned as a sphere with concentric shells of radius k (see Fig.
2) on which energy of the amount E(k) is continuously distributed.31 In our numerical simulations we
can only capture a part of the spectrum by choosing a discrete number of equidistant wavenumber shells
k1 , k1 + k, , k2 on which energy is distributed discretely among a limited number, Ns , of disturbance
modes for which k = kx2 + ky2 + kz2 . The scaling, f , must therefore be corrected since each mode on any
individual wavenumber shell contributes to the overall energy on that shell.
f=

2E(k)k
.
Ns

(18)

Finally, upon choosing the parameters L11 and T u, the inflow velocity- and vorticity fields can be completely
determinded. Figure 2 (right) presents instantaneous iso-surfaces for the spanwise disturbance velocity, w .
The iso-surfaces of w emphasize that the current method only introduces disturbances in the free stream,
but leaves the inside of the boundary layer almost undisturbed. Near the free stream the inflow disturbances
are ramped down to zero using an exponential ramping function. This has been found to be the best option
to yield numerically stable results.
E(k)

ky
-5/3

~x

k
2

Emax(u )
4

~x

kx

k1 k 1.157 k
max
L11

k2

kz

Figure 2. Left: Typical von K


arm
an energy spectrum (). The symbols ( ) mark the range of wavenumbers
realized in the numerical simulation. Center: schematic of the energy shell model. Energy is distributed
among a limited number of disturbance modes located on equidistant, spherical wavenumber shells. Right:
Instantaneous iso-surfaces of the spanwise disturbance velocity, w . The edge of the boundary layer of the
undisturbed flow is indicated by the boundary layer thickness, 99 .

III.

Validation

The original code has been developed and validated by Meitz32 for transition simulations of wall-bounded
shear flows. However, the reformulation of the governing equations and the associated changes in the code
made it necessary to confirm that correct simulation results are also obtained with the new code version. One
of the most common benchmark cases for computing the effect of FST on the laminar boundary layer and on
transition to turbulence are the experiments of Roach and Brierley.33 They reported detailed measurements
for nearly isotropic and homogeneous free-stream turbulence with intensities above 1% for a zero pressure
gradient boundary layer on a flat plate. A comparison between DNS results and measurements is presented
here for test case T3A of Roach and Brierley. For this case the free-stream turbulent intensity was 3% and
bypass transition occured.
6 of 21
American Institute of Aeronautics and Astronautics Paper 2010-4600

For the simulation setup we chose a rectangular domain similar to the one presented in Fig. 6 but with
zero pressure gradient. The flow parameters were U = 5.3 m/s, L = 0.1m, and = 1.5 105 m2 /s. The
integration domain started at x0 = 0.75 (Re1 = 280) and was extended to xmax = 15.15 (Re1 = 1258)
using 1601 equidistant grid points. A buffer domain technique (see Meitz32 ) is employed near the outflow
starting at xB = 13.89 to prevent reflections from the outflow boundary.
The height of the domain was ymax = 1.25 ( 5599,0 ) and exponential grid stretching34 was employed in
the wall-normal direction using 256 grid points. Typically, grid stretching is designed to cluster many points
near the wall. However, since the fluctuations in the free stream also need to be resolved sufficiently the grid
stretching was more benign than in transition simulations without a disturbed free stream. In the spanwise
direction, the domain size was Z = 0.7 ( 3199,0 ) and K = 64 cosine and K = 64 sine modes were used
to represent the flow field. The flow field was initialized using a Blasius similarity solution. Starting at
t = 0 the flow was subjected to a randomized inflow disturbance field which was generated for parameters
T u = 3.5% and L11 = 51,0 where 1,0 is the displacement thickness at the inflow. The disturbance field was
generated using a total of 600 continuous modes (20 wavenumber shells and Ns = 30 modes per shell). The
simulation was run for 4 flow-through times to allow the flow field to adjust to the turbulent free stream
and to flush out transients. Then time averages were taken over a period of 12 flow-through times. Finally,
turbulent statistical quantities were computed over a period of additional 6 flow-through times.
Figure 3 presents the time- and spanwise averaged skin-friction coefficient as a function of Reynolds
numbers Rex (based on downstream location, x) and Re (based on local momentum thickness, ). Also
shown are theoretical correlations for laminar and turbulent boundary layers.35 The agreement of both
curves with the T3A measurements is very good. Although no grid-resolution study is presented for the
current validation case, the fact that the curves from the DNS do not exhibit a severe overshoot when
passing the theoretical cf -curve for turbulent boundary layers, is an indicator for sufficient resolution in the
DNS.
0.008

0.008
Turbulent

Turbulent

0.006

cf

cf

0.006

0.004

0.004

0.002

0.002
Blasius

Blasius

0
0

110

210

310

410

510

0
0

200

400

600

800

1000

Re

Rex

Figure 3. Skin-friction coefficient, cf , plotted versus Reynolds number based on downstream location, Rex ,
(left) and momentum thickness, Re , (right). ( ) T3A experiments; () DNS; ( ) theoretical values for
laminar and turbulent boundary layers taken from White.35

In Fig. 4, time- and spanwise averaged streamwise velocity profiles plotted in wall coordinates are
compared to the experiment at various measurement locations. The differences in the results are minor
especially at the last station where the flow is fully turbulent and the profiles agree well with the log law.
Finally, we present characteristics of the turbulent fluctuations in the free stream. Figure 5 shows the
downstream decay of the relative turbulent intensities urms , vrms , and wrms which are defined as vi,rms =
(vi2 ). These quantities were measured at y = 2.599,B where 99,B is the boundary layer thickness at
the beginning of the buffer domain, xB . It can be seen that all velocity components of the fluctuating
disturbance field decay at the same rate and that the isotropic condition, i.e. urms vrms wrms , is
satisfied very well. The downstream decay of the turbulent intensity T u (see Eq. 17) is also shown in Fig.
5 and is found to agree very well with an analytical curve of the form T u = A(Rex + B)C where the decay
rate C = 0.74. This value is very close to values reported in the literature (C = 0.71,36 C = 0.5,37
C = 0.6210 ) for nearly isotropic turbulence generated by screens. The presented results for the validation
case are intentionally kept brief since the focus of the paper is on the influence of free-stream turbulence
on laminar separation and separation control. It should also be mentioned that the good agreement was
7 of 21
American Institute of Aeronautics and Astronautics Paper 2010-4600

30

Re=323
Rex=16.92e+4

10

10

10

10

y
30

10

10

10

10

y
30

Re=456
Rex=23.84e+4

10

10

Re=980
Rex=45.48e+4

20

20

Re=385
Rex=20.35e+4

20

20

30

10

10

10

10

10

10

10

10

10

10

Figure 4. Time- and spanwise averaged streamwise velocity profiles in wall coordinates taken at various
downstream locations (Rex , Re ). ( ) T3A experiments; () DNS based on Re ; (- - -) DNS based on Rex ;
( ) Spaldings law

0.05
urms

0.04

T3A

0.04

vrms

DNS

wrms

Tu=A(Rex+B)

0.03

0.03

Tu

velocity fluctuations

0.05

0.02

0.02

0.01

0.01

110 210 310 410 510

110 210 310 410 510

Rex

Rex

Figure 5. Downstream decay of turbulent fluctuations in the free stream at y = 2.599,B . Left: root-mean-square
values of the velocity fluctuations; right: turbulent intensity T u.

obtained for a fixed parameter L11 that was chosen as the lower limit of the integral length scale reported
in the T3A experiments (5mm to 30mm). Related studies of bypass transition show that the choice of L11
can have a significant effect on the transition process.38 Overall, the results in this section suggest that
the method for generating realistic turbulent free-stream conditions was implemented correctly and that
the updated version of the Navier-Stokes solver can be used for numerical investigations of transition in
wall-bounded shear flows.

IV.

Simulation Setup

The setup for the DNS with pressure gradient is motivated by wind-tunnel experiments conducted by
Gaster.2 In his experiments, Gaster mounted an inverted airfoil above a flat plate to accelerate and decelerate
the flow between the airfoil and the flat plate. Since circulation control was applied on the airfoil, flow
separation and reattachment occurred on the flat plate. By variation of the tunnel speed, Gaster generated
a series of short and long separation bubbles. In this paper we focus on a long bubble, Gasters case VI,

8 of 21
American Institute of Aeronautics and Astronautics Paper 2010-4600

series I.
For the numerical simulation, the governing equations are solved within a rectangular integration domain
shown schematically in Fig. 6. The velocity scale is chosen according to the tunnel speed in the experiment,
U = 6.64 m/s, and the length scale was L = 1in = 0.0254m. Also shown in Fig. 6 is the computational
grid. In the streamwise direction the baseline grid had 2001 equidistantly distributed points in the region
5.0 x 23.0. In order to save computational cost, simulations of active flow control or free-stream
turbulence are carried out in a shorter domain (1621 grid points). This is justified since the size of the
separation bubble is significantly reduced in these cases. For the wall-normal direction, 256 grid points
are used and exponential grid-stretching is applied. At the solid wall, y = 0, no-slip and no-penetration
boundary conditions are applied except for a narrow disturbance strip centered at x = 9 through which
disturbances can be introduced into the flow (see section V.C). The spanwise domain width is chosen as
Z = 2.0 ( 1999,0 ) and is resolved using K = 64 cosine and K = 64 sine modes or 200 collocation points,
respectively.

Figure 6. Simulation setup motivated by the experiments of Gaster.2 For the computational grid, every 10th
grid line is shown in both streamwise and wall-normal direction. Also shown are instantaneous gray-scale
contours of spanwise-averaged spanwise vorticity, z .

At t = 0 the flow field is initialized with the Blasius similarity solution. The pressure gradient along the
surface is enforced by choosing an appropriate boundary condition for the zeroth mode (spanwise average)
of the wall-normal velocity component, v, at the free-stream boundary. Typically a derivative condition of
the form v/y ymax = f (x) is employed to enforce the desired u/x in the free stream. However, for the
unforced case (T u = 0) such a condition was found to result in a separation bubble which continuously grows
until disturbances reach the free stream causing the termination of the simulation. We therefore decided to
use a more rigid Dirichlet condition v y = f(x) for which stable results can be obtained. However, also in
max
this case the separation bubble is always significantly longer than in the experiments. Satisfactory comparison
can only be obtained when some background noise is introduced into the separation bubble. In fact, it has
been a long-standing discussion whether the background disturbances, which originate from round-off and
truncation errors in a numerical simulation, are sufficient to reproduce realistic separation bubbles. For
high adverse pressure gradients and/or when flow instabilities are of absolute nature39 this can certainly be
the case. However, there are many cases reported in the literature28, 40 where appropriate forcing upstream
of separation is needed in order to match experimental data. As an alternative to introducing selective or
random disturbances inside the boundary layer, we propose to apply our numerical model for generating
isotropic grid turbulence since this approach is less artificial and is based on relevant physical mechanisms.

9 of 21
American Institute of Aeronautics and Astronautics Paper 2010-4600

V.
A.

Results

Uncontrolled, natural flow, T u = 0

In this section, we discuss results obtained for the uncontrolled flow without free-stream turbulence, i.e.
T u = 0. Instantaneous gray-scale contours of spanwise vorticity, z , obtained from 2-D and 3-D DNS are
presented in Fig. 7. For the 2-D case, the flow separates at xS = 10.5 followed by the shedding of strong
clockwise-rotating vortices (often referred to as rollers) which cause a reattachment of the flow (in the
mean) at xR = 13.1. These 2-D results predict a separation bubble that is much shorter than observed in
the experiments. Whereas flow transition is suppressed in a 2-D DNS, the 3-D flow can experience transition
to turbulence through a variety of physical mechanisms: primary and secondary instabilities, non-linear
generation of three-dimensional disturbance modes, etc. In the present case transition is observed to be
initiated in the separated shear layer above the wall. It is commonly argued that the break-up of the strong
spanwise vortices into small-scale turbulence causes the bubble to become longer since the turbulent mixing
provided by the small-scale structures is less effective than the 2-D rollers in transporting free-stream fluid
towards the wall. For the 3-D case, the mean separation and reattachment locations are xS = 10.4 and
xR = 15.5, respectively.

z
-0.5

0.0

0.5

*101
1.0

1.0
0.0

2.0

-1.0

5.0

6.0

7.0

8.0

9.0

10.0 11.0 12.0 13.0 14.0 15.0 16.0 17.0 18.0 19.0 20.0 21.0

1.0
0.0

2.0

5.0

6.0

7.0

8.0

9.0

10.0 11.0 12.0 13.0 14.0 15.0 16.0 17.0 18.0 19.0 20.0 21.0

x
Figure 7. Instantaneous gray-scale contours of the spanwise vorticity component, z , obtained from 2-D DNS
(top) and 3-D DNS (bottom). The 3-D result represents an average in spanwise direction.

Time- and spanwise averaged streamwise velocity profiles as well as streamlines are presented in Fig. 8.
The acceleration of the flow for x < 9 indicated by the decrease in boundary-layer thickness, can now be
better observed than in Fig. 7. The strong growth of the boundary layer downstream of reattachment and
the fuller velocity profiles show that the flow indeed transitions.
In order to assess the dynamic behavior of the separation bubble, time signals of the wall-normal disturbance velocity component, v = v v, are analyzed to detect dominant frequencies in the flow field. These
time signals are taken at the center of the domain, z = 0, and at a distance y = 1 (x) away from the wall
where 1 is the local, time- and spanwise averaged displacement thickness. This distance is chosen since
within the separated region, the displacement thickness is very close to the y-location for which the averaged
spanwise vorticity component, z , assumes a maximum in the wall-normal direction, 1 y
.
z,max
Gray-scale contours of ln |v | time-signals are presented in Fig. 9 in form of an x-t-diagram. Due to
the logarithmic scale, dark lines are an indication of regions where the disturbance velocity changes its
sign. For x < 12.5 downstream growth of v is observed with very little variation in time. Low-frequency
oscillations are a common characteristic of realistic separation bubbles caused by displacement effects. As
explained by Maucher et al.,41 the bubble grows and causes an acceleration of the outer flow which leads to a
reduction in the adverse pressure gradient. As a consequence of this, the bubble size decreases leading to an
increase in adverse pressure gradient. For x > 12.5 a variation of the velocity signal with a distinct timescale
becomes visible. These periodic contour variations are associated with the fundamental shedding cycle of
10 of 21
American Institute of Aeronautics and Astronautics Paper 2010-4600

1.0

0 0.5 1 0 0.5 1 0 0.5 1 0 0.5 1 0 0.5 1 0 0.5 1 0 0.5 1 0 0.5 1 0 0.5 1 0 0.5 1 0 0.5 1 0 0.5 1 0 0.5 1 0 0.5 1 0 0.5 1 0 0.5 1

0.5
99

0.0

11

10

12

14

13

15

16

17

18

19

20

21

16

17

18

19

20

21

1.0

0.5

0.0
5

10

11

12

13

14

15

x
Figure 8. Time- and spanwise averaged results for the unforced flow (T u = 0). Top: streamwise velocity
profiles. The dashed
line indicates the boundary layer thickness; bottom: streamlines, plotted in equidistant
`
contours of ln 104 + 1 .

the separation bubble. They show a positive slope in the x-t-diagram indicating that the disturbances are
travelling waves.

ln(|v'|)
-1.5

-1.0

-0.5

*101
0.0

0.0

10.0

20.0

30.0

-2.0

9.5

10.0

10.5

11.0

11.5

12.0

12.5

13.0

13.5

x
Figure 9. Gray-scale contours of the wall-normal disturbance velocity, v , as a function of downstream distance
and time.

Time signals of v and results of a spectral analysis for three downstream locations, x = 12.5, 13.0, 13.5,
are shown in Fig. 10. To determine the frequency spectrum associated with these time signals, we performed
a discrete Fourier transform (DFT) of the signals,
v =

1
N

N 1

vn eitn ,
n=0

Ev =

1
v v ,
2

(19)

where N is the number of samples in the time signal and = f /(2) is the frequency. We also performed
a so-called maximum entropy spectral analysis. The maximum entropy method (MEM) has its origin in
information theory where it is used for signal processing of data with high levels of broadband noise. MEM
estimates the spectral density of a discrete time signal by considering the most random, or least predictable,
process that has the same M + 1 autocorrelation coefficients as the given time series where M is called the
order of the method.42 Results of the spectral analysis for DFT and MEM using M = 10, 40 are shown
in Fig. 10. All spectra show the development of a distinct peak between f = 210Hz and f = 260Hz,
which corresponds to the shedding frequency of the separation bubble. The double-peak structure found
for x = 13.0 and x = 13.5 could be due to interference with the low-frequency bubble oscillation discussed
11 of 21
American Institute of Aeronautics and Astronautics Paper 2010-4600

earlier. The low-order MEM (M = 10) exhibits only one peak at x = 13.5 and we decided to chose this peak
with f = 240Hz as the fundamental frequency of the flow. At this x-location, a second peak with twice the
fundamental frequency can also be discerned from the plots of Fig. 10.
f=240Hz

0.02

-5

10

x=12.5
x=13.0
x=13.5

0.01

MEM (M=10)
MEM (M=40)
DFT

-7

10

Ev'

v'

-9

10

-11

10

-13

-0.01

10

-15

10

-0.02
0

0.05

0.1

10

100

t [s]

1000

f [Hz]

Figure 10. Left: Time signals of the wall-normal disturbance velocity, v , at y = 1 (x), z = 0 and various
downstream locations; right: spectral energy obtained using discrete Fourier Transform (DFT) and maximum
entropy method (MEM) of different order. The MEM results are scaled to have similar amplitudes as the
DFT results.

B.

Uncontrolled, natural flow, T u = 0

10.0

1.0
11.0

11.5

12.0

12.5

13.0

13.5

14.0

14.5

15.0

15.5

0.5

y
10.5

0.0

0.0

0.5

1.0

In his experiments, Gaster2 reported levels of turbulent free-stream fluctuations in the order of T u = 0.05%.
We used the method described previously (section II.A) to generate isotropic grid turbulence with T u =
0.05%, 0.5%, and 2.5%. For the turbulent integral length scale we chose L11 = 51,0 .
For the purpose of illustration we define a modified spanwise vorticity,
z = ln |z |. Instantaneous
contours of spanwise vorticity,
z , are plotted in Fig. 11. Upstream of transition, two different flow regions
can be identified for all cases: an outer region of turbulent fluctuations (except case T u = 0) and an inner
region of the laminar, separated boundary layer. The observation that the vortical free-stream disturbances
seem to be unable to enter the boundary layer is consistent with the so-called shear sheltering effect.43
This effect explains that free-stream disturbances are rapidly damped out inside a shear layer due to inviscid
shearing and viscous dissipation.

16.0

10.0

10.5

11.0

11.5

12.5

13.0

13.5

14.0

14.5

15.0

15.5

16.0

14.0

14.5

15.0

15.5

16.0

1.0
11.0

11.5

12.0

12.5

13.0

13.5

14.0

14.5

15.0

15.5

16.0

x
(c) T u = 0.5%

0.5

y
10.5

0.0

0.5
0.0

10.0

12.0

x
(b) T u = 0.05%

1.0

x
(a) T u = 0

10.0

10.5

11.0

11.5

12.0

12.5

13.0

13.5

x
(d) T u = 2.5%

Figure 11. Contours of 10.0 <


z < 10.0 where
z = ln |z | for various levels of free-stream turbulent intensity
T u. Dashed contour lines correspond to
z < 0.

The boundary layer, however, does not remain completely unaffected by free-stream disturbances. As
discussed in section I, it is well known that free-stream turbulence induces low-frequency u-velocity distortions
of large amplitude in the laminar region of the boundary layer - the so-called Klebanoff modes. Figure 12
presents instantaneous gray-scale contours of the streamwise disturbance velocity, u = u u
, plotted in
x-z-planes at distances y = 99 (x) and y = 1 (x) from the wall. Values for 99 (x) and 1 (x) (see Fig. 16) are
obtained from time- and spanwise averages of the flow field. From the u -velocity contours it can be seen that

12 of 21
American Institute of Aeronautics and Astronautics Paper 2010-4600

while the outer flow (y = 99 (x)) is characterized by the random fluctuations of the free stream, inside the
approaching laminar boundary layer (y = 1 (x)), the flow is organized by the presence of elongated, streaky
structures. Despite these spanwise variations, the separation bubble at low levels of FST (T u = 0.05%) still
seems to be governed by the inviscid shear-layer instability mechanism manifesting itself by the generation
of spanwise vortices seen in the region 12.0 < x < 14.0 in Fig. 12. For the case of high FST (T u = 2.5%),
however, the Klebanoff modes assume much greater amplitudes and as a direct consequence, the contours
at y = 1 (x) indicate that the separated shear layer loses its predominantly two-dimensional character and
experiences the onset of flow transition seemingly without the shedding of spanwise vortices. On the other
hand, the contours at y = 99 (x) give room for speculation that the reattaching flow might still be organized
by spanwise coherent flow structures.
Next, we performed a spectral analysis of v -velocity time signals taken at z = 0 and y = 1 (x) analogous
to the analysis presented in section V.A. Results are shown in Fig. 13, again in terms of gray-scale contours
of the reported time-signals and energy spectra at various downstream locations. Compared to the unforced
case (T u = 0), similar dynamic behavior is observed for the case of low FST (T u = 0.05%). Here, the
separated shear layer is relatively calm for x < 11.6 with only a small-amplitude variation in time followed
by the onset of vortex shedding. However, the downstream location for which this periodic time dependence
becomes visible in the gray-scale contours strongly varies between 11.2 < x < 12.0. As a result the doublepeak structure of the energy spectrum is even more pronounced in this case compared to the unforced case.

u'

(a) T u = 0.05%

-2.5

0.0

*10-4
5.0

2.5

y = 1

-1

-5.0

10

11

12

13

14

15

16

17

18

x
0
-1

y = 99

10

11

12

13

14

15

16

17

18

u'

(b) T u = 2.5%

-2.5

0.0

*10-2
5.0

2.5

y = 1

-1

-5.0

10

11

12

13

14

15

16

17

18

x
0
-1

y = 99

10

11

12

13

14

15

16

17

18

x
Figure 12. Instantaneous gray-scale contours of the streamwise disturbance velocity, u , at distances y = 1 (x)
and y = 99 (x) away from the wall.

13 of 21
American Institute of Aeronautics and Astronautics Paper 2010-4600

In turn, it can be speculated that this behavior is due to a stronger variation in the overall size of the
separation bubble. The strongest peak develops at f = 225Hz for the case of low FST. When the level of
FST increases, higher-amplitude (random) oscillations can already be observed in the upstream attached
part of the boundary layer. But even in these cases, the periodicity associated with vortex shedding is
distinguishable further downstream and the energy spectra for cases T u = 0.5% and T u = 2.5% show a
dominant frequency of f = 215Hz which is slightly lower than for the case T u = 0.
For all considered cases of added free-stream turbulent fluctuations, the observed dominant frequency is
close to the shedding frequency of the unforced separation bubble, f = 240 Hz. Therefore we conjecture
that the inviscid shear-layer instability mechanism must still be present in all cases and play an important
role in the transition process. To confirm this we looked more closely at the downstream development

ln(|v'|)
(a) T u = 0.05%

-2.0

-1.5

-1.0

-0.5

*101
0.0
f=225 Hz

-5

20.0

10

-7

10.0

Ev'

10

x=13.0

-9

10

x=12.5
x=12.0

-11

10

0.0

-13

10

9.6

10.0

10.4

10.8

11.2

11.6

12.0

12.4

10

12.8

100

1000

f [Hz]

x
(b) T u = 0.5%

f=215 Hz

-5

20.0

10

-7

10

10.0

Ev'

x=12.5

-9

10

x=12.0
x=11.5

-11

10

0.0

-13

10

9.6

10.0

10.4

10.8

11.2

11.6

12.0

12.4

10

12.8

100

1000

f [Hz]

x
(c) T u = 2.5%

f=215 Hz

-5

10

20.0

x=12.5

-7

10

10.0

Ev'

x=12.0

-9

10

x=11.5

-11

10

0.0

-13

10

9.6

10.0

10.4

10.8

11.2

11.6

12.0

12.4

12.8

10

100

1000

f [Hz]

Figure 13. Left: Spatial-temporal development of the wall-normal disturbance velocity component, v , at
locations (x, y = 1 , z = 0) for different FST levels; right: spectral energy obtained using DFT ( ) and MEM
of order M=40 () and M=10 ( ). The MEM results are scaled to have similar amplitudes as the DFT
results.

14 of 21
American Institute of Aeronautics and Astronautics Paper 2010-4600

of a two-dimensional u -velocity disturbance with the fundamental frequency. Results for this study are
presented in Fig 14. As the levels of FST increase, the disturbance amplitude in the approaching laminar
boundary layer reaches higher levels without necessarily experiencing significant downstream growth. In
the shear-layer, however, the disturbances grow exponentially (linear in the log-plot) until they saturate at
about the same amplitude level in each case. For all simulations with non-zero free-stream turbulence, the
onset of exponential growth is found at x 12.0 which is upstream of the location for which the undisturbed
case (T u = 0) shows exponential growth. For the undisturbed case we also performed a linear stability
analysis (solving equation 8a) using streamwise velocity profiles of the time- and spanwise averaged flow
field. Although linear theory is based on the parallel flow assumption it has been shown to be a reliable tool
for predicting disturbance growth especially in the upstream part of laminar separation bubbles.44
Sep.

10

-1

2.5

-2

10

y/1

Au'

10

-3

10

1.5

-4

10

-5

10

0.5

-6

10

10

12

0
0

14

0.5

u'/u' max

Figure 14. Left: downstream development of the Fourier amplitude for the disturbance velocity u (max.
over y) of the fundamental frequency f = 240 Hz; right: wall-normal amplitude distribution of the u -velocity
disturbance (normalized by its maximum) taken at the downstream location where Au = 1.5 102 . ( )
linear stability theory (LST), () T u = 0, ( ) T u = 0.05%, ( ) T u = 0.5%, ( ) T u = 2.5%.

Figure 14 also shows the wall-normal u -velocity amplitude distribution for each case. These profiles
are taken from the region of exponential growth for which the maximum of the disturbance amplitude has
reached Au = 1.5 102 . The shape of the amplitude distribution agrees very well with the u-velocity
eigenfunction obtained from linear stability theory (LST). Note that LST has only been applied for case
T u = 0 and that therefore the comparison of LST and DNS is only qualitative for the cases with T u = 0. In
summary, the good agreement between LST and DNS for the region of exponential growth confirms that the
2-D disturbances grow due to a linear instability mechanism (inviscid shear-layer instability) for all cases.
We also conclude that even for the highest level of FST considered in the present work, at least the primary
stage of the transition process is not completely bypassed.
Time- and spanwise averaged results of the different flow scenarios are compared in Figs. 15 and 16.
Wall-pressure- and skin-friction coefficients are plotted in Fig. 15. For x < 12.0 the curves are almost
identical. In particular, the separation location remains almost constant (xS 10.4), although a general
trend is observed in that the flow detaches slightly farther downstream when the FST increases. Locations
of separation and reattachment are summarized in Tab. 1. The agreement with the experimental results for
the wall-pressure coefficient is best for 0.05 < T u < 0.5 which indicates that it is necessary in this case to
introduce some level of realistic background disturbances in order to match the experimental measurements
(see discussion in section IV).

xS
xR

Tu = 0
10.40
15.54

T u = 0.05%
10.43
14.24

T u = 0.5%
10.47
13.67

T u = 2.5%
10.53
13.21

Table 1. Time-and spanwise averaged locations of separation, xS , and reattachment, xR .

15 of 21
American Institute of Aeronautics and Astronautics Paper 2010-4600

0.02

0.015

-0.25

0.01

cf

cp

0.25

-0.5

0.005

-0.75

-1

-0.005

-1.25

10

12

14

16

-0.01

18

10

12

14

16

18

Figure 15. Time- and spanwise averaged wall-pressure coefficient (left) and skin-friction coefficient (right) for
various levels of FST. ( ) experiment, () T u = 0, ( ) T u = 0.05%, ( ) T u = 0.5%, ( ) T u = 2.5%.

0.6

0.6

0.5

0.5

0.4

0.4

99

The boundary layer parameters 99 and 1 plotted in Fig. 16 show that the height of the separation
bubble is also significantly influenced by the free-stream turbulence. For example, the maximum displacement
thickness decreases 50% from case T u = 0 to case T u = 2.5%.

0.3

0.3

0.2

0.2

0.1

0.1

10

12

14

16

10

12

14

16

Figure 16. Time- and spanwise averaged boundary-layer thickness (left) and displacement thickness (right)
for various levels of FST. ( ) experiment, () T u = 0, ( ) T u = 0.05%, ( ) T u = 0.5%, ( ) T u = 2.5%.

C.

Controlled flow

In this section, we discuss simulations where high-amplitude, two-dimensional Tollmien-Schlichting waves


upstream of the separation location were generated by specifying a wall-normal velocity component, v, across
a narrow blowing and suction slot at the wall,
v(x, t) = vvs (x)cos(t) ,

(20)

where v, the amplitude, is chosen 5% of the free-stream velocity at the inflow. The angular frequency = 2f
is determined by the fundamental shedding frequency of the separation bubble (see section V.A) and vs (x)
is a shape function that is zero outside the suction/blowing slot and
x1 x xc :
x1 x xc :

1
7295 17014 + 9723 ,
48
1
vs () =
7295 17014 + 9723 ,
48
vs () =

x x1
xc x1
x2 x
=
x2 xc
=

(21a)
(21b)

within the slot, where x1 = 8.875, x2 = 9.125 and xc = (x1 + x2 )/2 mark the beginning, the end, and the
center of the disturbance slot, respectively. This forcing technique is deliberately chosen since it is known
that for a band of unstable frequencies, two-dimensional, harmonic blowing and suction can exploit the linear
shear-layer instability mechanism leading to exponential amplification of the initial disturbance wave.45 As
a consequence, separation is significantly reduced and in many cases transition can be delayed. In particular,
forcing with high-amplitude, two-dimensional TS waves has shown to delay the onset of a strong absolute
secondary instability for the LSB under investigation.46

16 of 21
American Institute of Aeronautics and Astronautics Paper 2010-4600

10.0

1.0
11.0

11.5

12.0

12.5

13.0

13.5

14.0

14.5

15.0

15.5

0.5

y
10.5

0.0

0.0

0.5

1.0

The effect of the applied forcing on the LSB is demonstrated in Fig. 17. Instantaneous contours of the
modified spanwise vorticity,
z , show the formation of spanwise rollers which remain close to the wall. As
discussed for the results of the 2-D DNS, these structures are very effective in transporting free-stream fluid
towards the wall and therefore reducing the size of the separation bubble. In fact, the results from the
3-D DNS presented in Fig. 17(a) appear as if they were obtained from a 2-D DNS since the flow remains
completely laminar in the entire integration domain. The formation of spanwise coherent structures can also
be observed in controlled wind- or water-tunnel experiments. However, relaminarization of the flow for long
downstream distances has never been observed experimentally. In the experiments, of course, the free stream
is not disturbance-free and the proper disturbance input for flow control is by no means trivial. Numerical
simulations including the effect of FST (Fig. 17(b)-(d)) show that, much like in the experiments, the flow
will eventually transition to turbulence in an environment with FST. Even for the smallest level of FST
considered in the current study (T u = 0.05%), the spanwise structures are seen to break up into smaller
scales at x > 14.5 initiating transition. When the levels of FST increase, the transition process starts further
upstream and for the highest FST level the spanwise structures appear to break up immediately.

16.0

10.0

10.5

11.0

11.5

12.5

13.0

13.5

14.0

14.5

15.0

15.5

16.0

14.0

14.5

15.0

15.5

16.0

1.0
11.0

11.5

12.0

12.5

13.0

13.5

14.0

14.5

15.0

15.5

16.0

x
(c) T u = 0.5%

0.5

y
10.5

0.0

0.5
0.0

10.0

12.0

x
(b) T u = 0.05%

1.0

x
(a) T u = 0

10.0

10.5

11.0

11.5

12.0

12.5

13.0

13.5

x
(d) T u = 2.5%

Figure 17. Contours of 7.5 <


z < 7.5 where
z = ln |z | for various levels of free-stream turbulent intensity
T u. Dashed contour lines correspond to
z < 0.

Note that from the contour lines in Fig. 17 it can not be concluded whether the transition process is
driven by the disturbances in the free stream or inside the shear layer. To answer this question we inspected
gray-scale contours of the streamwise disturbance velocity, u = u u
, plotted in x-z-planes at distances
y = 99 (x) and y = 1 (x) away from the wall for cases T u = 0.05% and T u = 2.5% (see Fig. 18). As in
Fig. 12, the ocurrence of the ubiquitous Klebanoff mode can be observed in the laminar boundary layer.
Downstream of the blowing- and suction slot (x > 9.0) the flow is dominated by the 2-D disturbance wave
introduced by the forcing. When the free-stream turbulent intensity is low (T u = 0.05%), these disturbances
remain two-dimensional over a large downstream distance before they start to break up at x 14. Spanwise
variations of the rollers are observed further upstream for the x-z-plane at y = 1 than for y = 99 . This is
also true for the case of high FST and, although not entirely conclusive, leads us to believe that the transition
process is most likely initiated by disturbances inside the boundary layer.
For high levels of FST the boundary layer streaks are stronger and immediately affect the 2-D waves as
can be inferred from the contours at y = 1 for case T u = 2.5%. However, the flow does not become turbulent
until x 12. At the edge of the boundary layer, y = 99 , and upstream of x 11 the flow field appears
to be dominated by the random fluctuations of the free stream rather than the 2-D wave. Downstream of
x > 11, and in contrast to the contours at y = 1 , a strong coherence in the spanwise direction can be seen.
This observation was already made in Fig. 11 and led us to speculate that the turbulent boundary layer is
organized by spanwise coherent flow structures.
Finally, in Fig. 19 we present a comparison of the time- and spanwise averaged wall-pressure and skinfriction coefficient for the uncontrolled and controlled flow cases. In all cases, the controlled flow shows a
significant reduction of separation. In the reattached, turbulent boundary layer, the skin-friction coefficient
is also reduced when flow control is employed. However, since the extent of the separated region is typically
small for applications with high levels of FST, the design requirements and the energy input required for the
control might outweigh the benefits of flow control.
In his experiments, Gaster also obtained surface pressure measurements for the case when the separation
bubble was supressed by tripping the laminar boundary layer to turbulence. These measurements are also

17 of 21
American Institute of Aeronautics and Astronautics Paper 2010-4600

u'

(a) T u = 0.05%

-2.5

0.0

*10-4
5.0

2.5

y = 1

-1

-5.0

10

11

12

13

14

15

16

17

18

y = 99

-1

10

11

12

13

14

15

16

17

18

x
u'

(b) T u = 2.5%

-2.5

0.0

*10-2
5.0

2.5

y = 1

-1

-5.0

10

11

12

13

14

15

16

17

18

y = 99

-1

10

11

12

13

14

15

16

17

18

x
Figure 18. Instantaneous gray-scale contours of the streamwise disturbance velocity, u , at distances y = 1 (x)
and y = 99 (x) away from the wall.

included in Fig. 19 and the comparison shows very good agreement for the controlled flow at T u = 2.5%.
Together with the flow visualization in Fig. 18(b) this raises the question if transition in this case might
occur independent of flow separation merely due to the interaction of the Klebanoff mode and the strong
2-D disturbances or if flow separation is necessary for the transition process.

VI.

Summary and Conclusions

Direct numerical simulations (DNS) of the Navier-Stokes equation were used to investigate the effect
of free-stream turbulence on laminar flow separation and separation control. For the numerical model of
realistic, isotropic grid turbulence, an approach based on the continuous modes of the Orr-Sommerfeld
spectrum (see Jacobs1 ) was implemented in our current Navier-Stokes solver. The method was validated
with experimental data available for bypass transition in a zero pressure gradient flat-plate boundary layer.
For DNS of a separating boundary layer, we investigated a laminar separation bubble on a flat plate. The
simulation setup was motivated by the wind-tunnel experiments by Gaster.2
Without adding any free-stream turbulent fluctuations, 2-D and 3-D DNS could not reproduce the wallpressure distribution reported in the experiments. The separation bubble was either too short (2-D) or
too long (3-D). However, when the free-stream turbulence intensity was increased to levels used in the
experiments (T u = 0.05%) good agreement was obtained using 3-D DNS. For the cases with FST we
studied three different turbulent intensities: T u = 0.05%, 0.5%, 2.5%. FST was found to induce elongated
streamwise streaks in the laminar boundary layer. The amplitude of these streaks was significantly larger for
higher levels of FST. For all cases, separation of the mean flow was not suppressed and transition was found

18 of 21
American Institute of Aeronautics and Astronautics Paper 2010-4600

(a) T u = 0

(b) T u = 0.05%

(c) T u = 0.5%

(d) T u = 2.5%

cp

0
-0.5
-1
8

12

16

12

16

12

16

12

16

12

16

12

16

12

16

12

16

0.02

cf

0.01
0
-0.01

Figure 19. Comparison of the wall-pressure coefficient for the uncontrolled flow ( ) and the controlled
flow () for various levels of FST. The (+)-symbols correspond to experimental data taken when the flow was
tripped to turbulence upstream of separation.

to occur in the separated shear layer above the wall. As the FST level increases, the extent and height of
the bubble were found to decrease. Using various methods of spectral analysis, we detected vortex shedding
of the shear layer with similar frequencies in all cases. Comparison to results obtained from linear stability
theory confirmed that even for the case of high FST (T u = 2.5%) the linear, inviscid shear-layer instability
mechanism plays an important role in the transition process
In additional simulations, we introduced two-dimensional, high-amplitude TS-waves in the attached laminar boundary layer prior to separation. From our experience, flow control using such highly-energetic
disturbance waves significantly reduces separation and can even delay transition until far downstream. For
increased levels of FST, the TS waves interacted with the three-dimensional streaks induced by the turbulent
free stream. Separation was still reduced in these cases but breakdown to turbulence was accelerated.

Acknowledgments
This work was supported by the Air Force Office of Scientific Research (AFOSR) under grant number
FA9550-09-1-0214, with Douglas R. Smith serving as the program manager. Computer time was provided by
the US Army Space and Missile Defense Command (SMDC), the Navy DoD Supercomputing Resource Center
(DSRC) and the Arctic Region Supercomputing Center (ARSC) under challenge project AFOSR13312C4A.

References
1 Jacobs,

R. G., Bypass Transition Phenomena studied by Computer Simulation, Ph.D. thesis, Stanford University, 2000.
M., The structure and behaviour of laminar separation bubbles, AGARD CP 4 , 1966, pp. 813854.
3 Hourmouziadis, J., Das DFG Verbundvorhaben Instation
are Str
omung in Turbomaschinen, Tech. rep., DGLR-JT2000030, 2000.
4 Schubauer, G. B. and Skramstad, H. K., Laminar boundary layer oscillations and transition on a flat plate, Tech. Rep.
909, NASA, 1948.
5 Dryden, H. L., Air Flow in the boundary layer near a Plate, 1936, NACA Rep. 562.
2 Gaster,

19 of 21
American Institute of Aeronautics and Astronautics Paper 2010-4600

6 Klebanoff, P. S. and Tidstrom, K. D., Evolution of amplified waves leading to transition in a boundary layer with zero
pressure gradient, 1959, NASA TN D 195.
7 Klebanoff, P., Effect of freestream turbulence on the laminar boundary layer, Bull. Am. Phys. Soc, Vol. 10, 1971,
pp. No. 11.
8 Kendall, J. M., Experimental study of disturbances produced in a pre-transitional laminar boundary layer by weak
freestream turbulence, AIAA Paper , Vol. 85-1695, 1985.
9 Kendall, J. M., Boundary layer receptivity to freestream turbulence, AIAA Paper , Vol. 90-1504, 1990.
10 Westin, K. J. A., Boiko, A. V., Klingmann, B., B. G., Kozlov, V., V., and Alfredsson, P. H., Experiments in a boundary
layer subjected to free-stream turbulence. Part 1. Boundary Layer structure and receptivity. J. Fluid Mech., Vol. 281, 1994,
pp. 193218.
11 Mayle, R. E., The Role of Laminar-Turbulent Transition in Gas Turbine Engines, ASME J. Turbomach., Vol. 113,
1991, pp. 509537.
12 Gault, D., An experimental investigation of regions of separated laminar flow, Tech. rep., 1955, NACA TN 3505.
13 Volino, R. J., Separated Flow Transition under Simulated Low-Pressure Turbine Airfoil Conditions Part1: Mean Flow
and Turbulence Statistics, ASME J. Turbomach., Vol. 124, 2002, pp. 645655.
14 Dong, Y. and Cumpsty, N. A., Compressor Blade Boundary Layer: Part2 Measurements with incident wakes, ASME
J. Turbomach., Vol. 112, 1990, pp. 231241.
15 Sohn, K. and Reshotko, E., Experimental Study of Boundary Layer Transition with Elevated Freestream Turbulence on
a Heated Flat Plate, 1991, NASA CR-187068.
16 H
aggmark, C., Investigations of disturbances developing in a laminar separation bubble flow, Ph.D. thesis, Royal Institute
of Technology, Stockholm, 1987.
17 Morkovin, M. V., On the Question of Instabilities Upstream of Cylindrical Bodies, Tech. rep., NASA, 1979.
18 Durbin, P. and Wu, X., Transition beneath vortical disturbances, Ann. Rev. Fluid Mech., Vol. 39, 2007, pp. 107128.
19 Andersson, P., Berggren, M., and Henningson, D., Optimal Disturbances and Bypass transition in Boundary Layers,
Phys. Fluids, Vol. 11, 1999, pp. 134150.
20 Luchini, P., Reynolds number independent Instability of the Boundary Layer over a flat surface, Part2: Optimal
Perturbations, J. Fluid Mech., Vol. 404, 2000, pp. 289309.
21 Wu, X., Jacobs, R., Hunt, J., and Durbin, P. A., Simulation of boundary layer transition induced by periodically passing
wakes, J. Fluid Mech., Vol. 398, 1999, pp. 109153.
22 Wu, X. and Durbin, P. A., Evidence of longitudinal vortices evolved from distorted wakes in a turbine passage, J. Fluid
Mech., Vol. 446, 2001, pp. 199228.
23 Reshotko, E., Transient Growth: A factor in bypass transition, Phys. Fluids, Vol. 13, No. 5, 2001, pp. 10671075.
24 Reshotko, E. and Tumin, A., Investigation of the Role of Transient Growth in Roughness-Induced Transition, AIAA
Paper , Vol. 02-2850, 2002.
25 Fasel, H. F., Numerical investigation of the interaction of the Klebanoff-mode with a Tollmien-Schlichting wave, J.
Fluid Mech., Vol. 450, 2002, pp. 133.
26 Rist, U. and Fasel, H., Direct numerical simulation of controlled transition in a flat-plate boundary layer, J. Fluid
Mech., Vol. 298, 1995, pp. 211248.
27 Fasel, H. F., Recent developments in the numerical solution to the Navier-Stokes and hydrodynamic stability problems,
Computation Fluid Dynamics, 1980, pp. 167280.
28 Marxen, O., Numerical Studies of Physical Effects Related to the Controlled Transition Process in Laminar Separation
Bubbles, Ph.D. thesis, Universit
at Stuttgart, 2005.
29 Davies, C. and Carpenter, P. W., A Novel Velocity-Vorticity Formulation of the Navier-Stokes Equations with Applications to Boundary Layer Disturbance Evolution, J. Comp. Phys., Vol. 172, Sept. 2001, pp. 119165.
30 Grosch, C. E. and Salwen, H., The continuous spectrum of the Orr-Sommerfeld equation. Part 1. The spectrum and the
eigenfunctions, J. Fluid Mech., Vol. 68, 1978, pp. 3354.
31 Brandt, L., Schlatter, P., and Henningson, D. S., Transition in boundary layers subject to free-stream turbulence, J.
Fluid Mech., Vol. 517, 2004, pp. 167198.
32 Meitz, H., Numerical Investigation of Suction in a Transitional Flat-Plate Boundary Layer, Ph.D. thesis, University of
Arizona, 1996.
33 Roach, P. E. and Brierley, D. H., The influence of a turbulent free-stream on zero pressure gradient transitional boundary
layer development Part1: Test cases T3A and T3B, ERCOFTAC, 1992, pp. 319347.
34 Meitz, H. and Fasel, H. F., A compact-difference scheme for the Navier-Stokes Equations in Vorticity-Velocity Formulation, J. Comp. Phys., Vol. 157, 2000, pp. 371403.
35 White, F. M., Viscous Fluid Flow , McGraw-Hill, 2nd ed., 1991.
36 Baines, W. D. and Petersen, E. G., An investigation of flow through screens, Trans. ASME , Vol. 73, 1951, pp. 467480.
37 Fasel, H. F., Turbulence reduction by screens, J. Fluid Mech., Vol. 197, 1976, pp. 139155.
38 Ovchinnikov, V., Choudhari, M. M., and Piomelli, U., Numerical simulations of boundary-layer bypass transition due
to high-amplitude free-stream turbulence, J. Fluid Mech., Vol. 613, 2008, pp. 135169.
39 Spalart, P. R. and Strelets, M. K., Mechanisms of transition and heat transfer in a separation bubble, J. Fluid Mech.,
Vol. 403, 2000, pp. 329349.
40 Baragona, M., Unsteady Characteristics of laminar separation bubbles. An experimental and numerical investigation,
Ph.D. thesis, Technische Universiteit Delft, 2004.
41 Maucher, U., Rist, U., and Wagner, S., Direct Numerical Simulation of Airfoil Separation Bubbles, S. Wagner, E.H.
Hirschel, J. Priaux, R. Piva (Eds.): Computational Fluid Dynamics 94 , John Wiley & Sons Ltd., 1994, pp. 471477.

20 of 21
American Institute of Aeronautics and Astronautics Paper 2010-4600

42 Ghil, M., Allen, M. R., Dettinger, M. D., Ide, K., Kondrashov, D., Mann, M. E., Robertson, A. W., Saunders, A., Tian,
Y., Varadi, F., and Yiou, P., Advanced spectral methods for climatic time series, Rev. Geophys., Vol. 40(1), 2002.
43 Hunt, J. C. R., Durbin, P. A., Kevlahan, N. K.-R., and Fernando, H. J. S., Non-local effects of shear in turbulent flows,
Sixth European Turbulence Conference, Lausanne, Switzerland, 1996.
44 Marxen, O., Rist, U., and Wagner, S., Effect of spanwise-modulated disturbances on transition in a separated boundary
layer, AIAA J., Vol. 42, 2004, pp. 937944.
45 Gross, A., Balzer, W., and Fasel, H. F., Numerical Investigation of Low-Pressure Turbine Flow Control (invited), AIAA
Paper , Vol. 08-4159, 2008.
46 Embacher, M., Numerical Investigation of Secondary Absolute Instability in Laminar Separation Bubbles, Masters thesis,
University of Arizona, 2006.

21 of 21
American Institute of Aeronautics and Astronautics Paper 2010-4600

You might also like