You are on page 1of 5

A new mechanism for rapid transition involving a pair of oblique waves

P. J. Schmid and D. S. Henningson


Citation: Physics of Fluids A 4, 1986 (1992); doi: 10.1063/1.858367
View online: http://dx.doi.org/10.1063/1.858367
View Table of Contents: http://scitation.aip.org/content/aip/journal/pofa/4/9?ver=pdfcov
Published by the AIP Publishing
Articles you may be interested in
Propagation and oblique collision of electrostatic solitary waves in quantum pair-plasmas
Phys. Plasmas 17, 082317 (2010); 10.1063/1.3480307
Interaction of an oblique shock wave with a pair of parallel vortices: Shock dynamics and mechanism of
sound generation
Phys. Fluids 18, 126101 (2006); 10.1063/1.2391806
Scattering of a horizontally polarized oblique wave by a pair of surfacebreaking cracks
J. Acoust. Soc. Am. 100, 2937 (1996); 10.1121/1.417104
A New Mechanism for Oblique Wave Resonance
Phys. Fluids 6, S8 (1994); 10.1063/1.4739106
New Magnetic Oxides Involving Divalent Europium and Transitional Metal Ions
AIP Conf. Proc. 18, 749 (1974); 10.1063/1.3141819

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Download to IP: 177.136.148.4 On: Sun, 18
Sep 2016 05:26:22

A new mechanism for rapid transition

involving a pair of oblique waves

P. J. Schmid and D. S. Henningson


Department of Mathematics, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139

(Received 27 December 1991; accepted 3 March 1992)


A pair of small but finite-amplitude oblique waves were used as initial condition in numerical
simulations of transition to turbulent flow. A rapid energy growth to a transition peak
occurred. Although this transition scenario is triggered by a nonlinear interaction, the
dominant mechanism yielding substantial disturbance amplitudes was found to be
linear. A number of simulations were made that showed that the overall features of this
transition scenario remains the same for a range of parameter values.

I. INTRODUCTION

Calculations of the stability of a flow to perturbations


are often based on linear stability theory, which for incompressible parallel shear flows predicts that the disturbance
that first experiences exponential growth is a twodimensional one. The two-dimensional waves, in turn, become unstable to three-dimensional waves, first seen experimentally by Klebanoff et al. r and theoretically by Orszag
and Patera and Herbert.3 For channel flow, this secondary
instability scenario operates down to a Reynolds number of
about 1000 (based on the centerline velocity and the channel half-height), if one artificially introduces a largeamplitude two-dimensional wave. The total disturbance
energy growth is quite slow, however, and the initial energy required to obtain growth rather large.
Morkovin4 proposed the concept of bypass transition as
an alternative route to turbulence, circumventing the secondary instability process. This transition scenario occurs
on a fast time scale and allows for a disturbance environment that is initially three dimensional.
Bypass transition has been studied by Henningson et
al. 5 for localized disturbances in plane Poiseuille flow and
by Lu and Henningson who idealized those disturbances
by considering the interaction of oblique modes corresponding to the peaks in the energy spectrum of the localized disturbances. Both studies found a rapid increase in
disturbance energy mostly associated with structures elongated in the streamwise direction, which in the latter investigation, was found to result in sustained turbulentlike
fluctuations.
The present work uses the same type of initial condition as in Ref. 6, i.e., a symmetric pair of oblique OrrSommerfeld (OS) modes. Here, we will address their nonlinear interaction and subsequent breakdown as well as the
energy transfer mechanisms causing the rapid growth.
II. NUMERICAL

OBSERVATIONS

Direct numerical simulations of the incompressible


Navier-Stokes equations for the channel flow geometry are
used to calculate the evolution of the oblique waves. The
code is based on spectral methods and is the same as used
in Ref. 5. Spanwise symmetry has been imposed, the calculations presented have been carried out with constant
mass flux and all flow quantities have been nondimension1986

Phys. fluids A 4 (9), September

1992

alized by the laminar centerline velocity (U,,) and the


channel half-height (h). For Re= U,-,h/v=
1500, 96 X 97
X 192 spectral modes have been used (reduced to half by
the spanwise symmetry) and, for Re=2000, the number of
modes was 128X97X 192.
In order to describe the mechanism involved in the
initial stages of transition from a symmetric pair of two
oblique waves, we analyze a typical case at Re= 1500. We
choose the two oblique waves to have the wave-number
vector (or,&) = ( 1, f 1) with the two components representing streamwise and spanwise wave numbers, respectively, and an initial maximum amplitude of Ac=0.05 in
the normal velocity. Note that all of the wave numbers that
will be excited during the development of the disturbance
are multiples of the initial ones, i.e., (o,,$,J = ( mal,&>,
where m and n are integers.
The flow field evolving from the oblique waves is characterized by a rapid development into a turbulentlike state.
Figure 1 shows that the total disturbance energy as a function of time has a pronounced peak at a time of about 28
before it levels off into a quasiequilibrium state. The disturbance energy without the mean flow component reaches
a peak earlier, although the maximum is lower. The substantial difference between the two curves anticipates a
strong involvement of the mean flow component during the
transition process. This simulation has been continued further in time in order to ensure that the disturbances are
sustained. The velocity fields past the transition peak show
sharp shear layers and streaky structures close to the wall,
confirming that a turbulentlike state has been reached.
Ill. ANALYSIS

OF ENERGY

TRANSFER

To better understand the rapid transition found for


this disturbance, we now consider the evolution of the energy in wave-number space. In Fig. 2(b) the energy in the
initially excited mode is displayed and Fig. 2(a) shows the
strong mean flow distortion. Figures 2(c) and 2(d) show
the energy in the wave numbers corresponding to the first
and second generation of nonlinear interactions, respectively. Note that the modes with low streamwise wave
numbers dominate in both cases. The reason for this wavenumber selection and the resulting anisotropic energy
propagation in wave-number space can be found by considering the Fourier-transformed equations governing the

0899-8213/92/091986-04$04.00

@ 1992 American Institute of Physics

1986

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Download to IP: 177.136.148.4 On: Sun, 18
Sep 2016 05:26:22

6.0 -

0.0 I

0.

10.

20.

30.

40.

50.

0.

10.

20.

30.

40.

50.

0.

10.

20.

30.

40.

50.

0.

10.

20.

30.

40.

50.

(a)
FIG. 1. Time evolution of the total disturbance energy (solid) and the
disturbance energy with the mean (0,O) component subtracted (dashed).

disturbance. For the wave number (a,,&),


the equations
for the normal velocity (3 and normal vorticity (?j), can
be written
$ Mq,n,=Lq,n,+

W,ai$

k+i=m
l+j=n

where qmn= ( cm,,;l^,,) T and


(b)

M=(-gg.

;)

M-lL=(l$s,

$,).

(1)

Here, v2 is the Fourier-transformed Laplacian, Tos and


?ZsQ are the Orr-Sommerfeld (OS) and Squire (SQ) operators, respectively, and N( .,a) is the nonlinear interaction term. For the explicit form of the above operators, see
Henningson and Schmid. Using this formalism, the evolution equation for the spectral energy
Em=&

I 1 dt,Mq,n

can be written
-$E&

8
,?lIl

c
k+i=m
Ifj=ll

1
-1

&$%W&j)dY

(2)

where pmn=c& +& This expression shows a production


term due to a linear mechanism and another due to nonlinear wave triad interactions. Since the governing linear
operator (M-IL)
is non-normal (a non-normal operator
has nonorthogonal eigenfunctions), the solutions to the
corresponding linear initial value problem can experience
large transient growth, even if all of the associated eigenvalues are damped (Reddy et al. *). This can occur when
the initial disturbance is comprised, not of a single mode,
but of a sum of nonorthogonal eigenfunctions that, to a
large extent, cancel each other out. From a physical point
of view, on the other hand, most of the growth associated
with the linear operator can be related to the vortex tilting

2.5

0.0
(d)

FIG. 2. Energy in various wave-number components. (a) The mean


distortion (0,O). (b) The initial wave (1,l). (c) The first generation of
nonlinear interactions, solid: (0,2), dash: (2,2), dot: (2,O). (d) The second generation of nonlinear interactions, solid: (1,3), dash: (3,3), dot:
(3,l). Note the difference in vertical scales.

1987
Phys. Fluids A, Vol. 4, No. 9, September 1992
P. J. Schmid and D. S. Henningson
1987
Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Download to IP: 177.136.148.4 On: Sun, 18
Sep 2016 05:26:22

term in the normal vorticity equation, here represented by


i&U in the off-diagonal term of the second matrix operator in ( 1). Naturally, the nonlinear terms may also contribute to the growth of a particular Fourier component,
although they drop out in the evolution equation for the
total energy obtained by summing (2) over all wave numbers.
Figure 3 displays the most dominant energy transfer
terms for the wave numbers found to attain the largest
amplitudes during the initial phases of the transition process. In all cases, we observe that the linear growth mechanism dominates. In Fig. 3 (a) the initial linear energy
transfer into the mode ( 1,l) is due to the omission of the
normal vorticity component associated with the initially
excited oblique OS modes. For such an initial condition,
the vortex tilting term acts as a forcing on the normal
vorticity, resulting in transient growth. Figure 3 (b) shows
that, after a short period of nonlinear energy transfer from
the ( 1,l) mode, the linear mechanism takes over. In this
case, it can be related to a near-degeneracy between OS and
SQ modes along the spanwise wave-number axis.7 This
near-degeneracy is a manifestation of the non-normality
discussed earlier and it is the main reason for the preferred
excitation of modes with low streamwise wave numbers.
The linear transfer into the (0,2) mode is counterbalanced
by a strong nonlinear interaction with the mean flow. The
results of Fig. 3(c) show the same qualitative behavior,
although the nonlinear generation phase is shorter, as can
be seen in Fig. 3 (d). Thus the nonlinear interactions redistribute the energy in wave-number space, while a major
part of the growth of the most excited wave numbers is due
to a linear mechanism. When the transition peak is approached, however, a number of additional Fourier components are excited and the dominance of the linear transfer mechanism diminishes.
IV. PARAMETER

STUDY

In an attempt to quantify the effect of different parameters on the described transition scenario, the following
characteristic quantities have been introduced: (i) the peak
time P such that (&Y/C&) (P) =0, (ii) the energy amplification factor EpE& and (iii) the equivalent growth rate
(T = ( l/P)ln(E~E~),
where Eh denotes the initial disturbance energy and E& stands for the corresponding energy
peak (excluding the mean flow component, dashed line in
Fig. 1).
Table I displays the three quantities for different combinations of physical parameters. The results show a weak
dependence on the Reynolds number. By increasing the
initial amplitude, the time to reach the energy peak as well
as the energy amplification factor decrease considerably,
resulting in a small change in the equivalent growth rate.
The effect of increasing the wave angle is mainly to decrease the peak time, whereas the energy amplification factor is observed to be rather insensitive to changes in the
wave angle. This implies an increase of the equivalent
growth rates. It is interesting to note that both t* and
EKES appear to be inversely proportional to the initial
1988

Phys. Fluids A, Vol. 4, No. 9, September

1992

x 10-s
:;T----7

-1.01

(a) 0.

5.

10.

1.5. t 20.

25.

30.

x 10
2-o1
1.5 -

---=----:_;-_-...
. ...l,

-0.5 -1.0 5.

I
10.

I
15.

-1.01

(=) 0.

5.

10.

15.

(b) 0.

____

,/
.\\
I-..___/ I
t 20.
25.

,/- __--30.

x 10
2.01

,
t 20.

25.

30.

x 10-5

FIG. 3. Temporal evolution of energy transfer rates. (a) Into ( 1,l) : sum
of interaction with (2,O) + (- 1,l) and (0,2) + (I,- 1) (chain-dashed),
interaction with (1,3) + (0,-2)
(chain-dotted). (b) Into (0,2): interactions with (l,l)+(-1,l)
(chain-dashed), with (1,3)+(-1,-l)
(chain-dotted). (c) Into (1,3): interaction with (0,2) + (1,l) (chaindashed). (d) Detailed view of initial stage for (1,3): interaction with
(2,2) + ( - 1,l) (dotted). The linear transfer is displayed by solid lines
and the mean Bow interaction by dashed lines.
P. J. Schmid and D. S. Henningson

1988

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Download to IP: 177.136.148.4 On: Sun, 18
Sep 2016 05:26:22

TABLE I. Parameter study of the transition process. Here, A, is the


maximum initial amplitude in the normal velocity and the angle refers to
arctan pi/a,, where a,= 1.
P
Re
1500
2mo

Re
1500
2ooo

A0
0.05
0.08
0.05
0.08
Ao

45

53

61"

20.8
15.0
20.0
14.5

18.2
14.0
16.9
13.6

15.7
7.9
14.8
7.7

53

61"

EPE;
45

0.05
0.08
0.05
0.08

10.8
5.3
11.8
5.6

Ao

45

11.4
5.3
11.9
5.5

9.2
4.8
9.8
4.9

*
Re
1500
2OCQ

0.05
0.08
0.05
0.08

0.114
0.111
0.124
0.120

53
0.134
0.119
0.146
0.130

61"
0.142
0.200
0.154
0.207

amplitude of the oblique waves. However, the number of


simulated cases are not large enough to be able to draw any
definite conclusions regarding such a scaling.
For the Reynolds numbers considered, the initial energy for &=0.08
(corresponding to Eh z 8 X 10M3) is
approximately the same as the lowest needed to obtain
growth for secondary instability disturbances.2 The largest
exponential growth rates obtained for those types of disturbances are about the same as the values of the equivalent exponential growth rates found for the interacting oblique waves. Note, however, that the growth rates in the
secondary instability scenario refer to the small threedimensional disturbances superimposed on the decaying
large-amplitude two-dimensional (2-D) wave, whereas the
present growth rates refer to the total disturbance energy
E. Thus, at the time of the transition peak in the present
scenario, the total disturbance energy may still be decaying
for a typical subcritical secondary instability disturbance.
V. DISCUSSION

In summary, a transition scenario has been found that


will occur at a much faster time scale than the one for
secondary instability, starting with equivalent initial disturbance energy. This is achieved by bypassing the rather
slow secondary instability stage, directly starting with a
pair of finite-amplitude oblique waves. It is plausible to
assume, however, that once the secondary instability has
produced a pair of finite-amplitude oblique waves, their
interaction will trigger the same linear mechanism described in the present paper.
The development of the oblique waves is dominated by
a preferred spreading of spectral energy into modes with
low streamwise wave numbers. The major mechanism
yielding substantial growth of perturbation energy has
been identified to be a linear one, balanced by a mean flow

interaction. The nonlinear terms were mainly responsible


for an initial transfer of energy into modes that then grow
by the linear mechanism. When the transition peak is approached, however, the dominance of the linear transfer
mechanism diminishes. A parameter study revealed that
this transition scenario is speeded up by an increase in
amplitude or wave angle but is rather insensitive to
changes in the Reynolds number. For all cases studied, the
qualitative features of the scenario remained the same.
The prominence of the (0,2) component (often called
the streamwise vortex mode) seen in the present transition
scenario has also been noted in compressible simulations of
transition on a cone, in simulations of transition in compressible boundary layers and in simulations of incompressible boundary-layer transition using the parabolized
stability equations. It is interesting to note that Pruett
and Zang started with a secondary instability type disturbance which in turn generated a dominating (0,2) component. This suggests that the linear mechanism described
here is a generic one and is operating in a wide variety of
flow situations and is not only relevant for the particular
transition scenario investigated in this study.

ACKNOWLEDGMENTS

This work was started at NASA Langley Research


Center, Hampton, VA (LaRC) while the authors attended
the 1991 ICASE workshop on transition to turbulence.
Computer time was provided by LaRC and NCSA, Urbana, IL.

P. S. Klebanoff, K. D. Tidstrom, and L. M. Sargent, The threedimensional nature of boundary layer instability, J. Fluid Mech. 12, 1
(1962).
*S. A. Orszag and A. T. Patera, Secondary instability of wall-bounded
shear flows, J. Fluid Mech. 128, 347 (1983).
T. Herbert, Modes of secondary instability in plane Poiseuille flow, in
Turbulence and Chaotic Phenomena in Fluids, edited by T. Tatsumi
(Elsevier, New York, 1984), pp. 53-58.
4M. V. Morkovin, The many faces of transition, in Viscous Drag Reduction, edited by C. S. Wells (Plenum, New York, 1969).
5D. S. Henningson, A. V. Johansson, and A. Lundbladh, On the evolution of localized disturbances in laminar shear flows, in LaminarTurbulent Transition, edited by D. Arral and R. Michel (SpringerVerlag, New York, 1990), pp. 279-284.
6Q. Lu and D. S. Henningson, Subcritical transition in plane Poiseuille
flow, Bull. Am. Phys. Sot. 35, 2288 (1990).
D. S. Henningson and P. J. Schmid, Vector eigenfunction expansions
for plane channel flows, Stud. Appl. Math. 87, 15 (1992).
*S. C. Reddy, P. J. Schmid, and D. S. Henningson, Pseudospectra of the
OrrSommerfeld
operator, to appear in SIAM J. Appl. Math.
C. D. Pruett and T. A. Zang, Direct numerical simulations of laminar
breakdown in high speed axi-symmetric boundary layers, AL4A Paper
No. 92-0742, 1992.
OH. Fasel and A. Thumm, Direct numerical simulation of threedimensional breakdown in supersonic boundary layer transition, Bull.
Am. Phys. Sot. 36, 2701 (1991).
IF. P. Bertolotti and T. Herbert, Fast simulation of boundary-layer
transition with PSE, Bull. Am. Phys. Sot. 35, 2230 (1990).

1989
P.J.Schmid
and D.S.Henningson
Phys. Fluids A, Vol. 4, No. 9, September 1992
1989
Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Download to IP: 177.136.148.4 On: Sun, 18
Sep 2016 05:26:22

You might also like