You are on page 1of 17

Theoret. Comput.

Fluid Dynamics (2001) 14: 359375


Theoretical andComputational
FluidDynamics
Springer-Verlag 2001
Prehistory of Instability in a Hypersonic Boundary Layer

Alexander V. Fedorov and Andrew P. Khokhlov


Moscow Institute of Physics and Technology,
16 Gagarin Street, Zhukovski, Moscow Region 140180, Russia
afedorov@pt.comcor.ru
Communicated by M.Y. Hussaini
Received 17 July 2000 and accepted 23 March 2001
Abstract. The initial phase of hypersonic boundary-layer transition comprising excitation of boundary-
layer modes and their downstream evolution from receptivity regions to the unstable region (instability
prehistory problem) is considered. The disturbance spectrum reveals the following features: (1) the rst
and second modes are synchronized with acoustic waves near the leading edge; (2) further downstream,
the rst mode is synchronized with entropy and vorticity waves; (3) near the lower neutral branch of the
Mack second mode, the rst mode is synchronized with the second mode. Disturbance behavior in Re-
gions (2) and (3) is studied using the multiple-mode method accounting for interaction between modes
due to mean-ow nonparallel effects. Analysis of the disturbance behavior in Region (3) provides the in-
termodal exchange rule coupling input and output amplitudes of the rst and second modes. It is shown
that Region (3) includes branch points at which disturbance group velocity and amplitude are singular.
These singularities can cause difculties in stability analyses. In Region (2), vorticity/entropy waves are
partially swallowed by the boundary layer. They may effectively generate the Mack second mode near its
lower neutral branch.
Nomenclature
A eigenvector of direct problem
B eigenvector of adjoint problem
c amplitude coefcient or phase speed
F =
e
/U
e
2
frequency parameter
L streamwise length
M Mach number
P mean-ow pressure
p pressure disturbance
R =
_
U

e
x

e
Reynolds number
T mean-ow temperature
t time
U, V streamwise and vertical velocity of the mean ow
u, v disturbance of x- and y-velocity components

This work was sponsored by the European Ofce of Aerospace Research and Development of the Department of the Air Force under
Contract F61708-96-W0196.
359
360 A.V. Fedorov and A.P. Khokhlov
x = x

/L

streamwise coordinate
y = y

/ coordinate normal to the wall


wave number
=
_
L

e
/U

e
boundary-layer scale

boundary-layer displacement thickness


viscosity
kinematic viscosity
temperature disturbance
Pr Prandtl number
angular frequency
Subscripts
ad adiabatic wall
asterisk synchronization of entropy/vorticity waves with Mode 1
b branch point
e upper boundary layer edge or entropy wave
i imaginary part
r real part
vorticity wave
w wall
Superscripts
bar complex conjugate
asterisk dimensional
1. Introduction
For free ight and quiet wind-tunnel conditions, freestream disturbances and perturbations generated by
hypersonic body surfaces are normally small in a high-frequency band relevant to laminar-turbulent transi-
tion. In this case the transition process comprises excitation of unstable normal modes (receptivity problem),
their downstream amplication (stability problem), and nonlinear breakdown (Malik et al., 1990; Morkovin,
1969; Reshotko, 1976). Linear stability theory (LST) helps to identify unstable modes and predict their
downstream growth (Mack, 1984; Malik, 1989). LST is well developed for subsonic and moderate super-
sonic boundary layers (Reshotko, 1994). Its results are consistent with many experiments (Kachanov et al.,
1982; Gaponov and Maslov, 1980; Kosinov et al., 1990). Good progress has been made in the receptivity
of these ows (Zhigulev and Tumin, 1987; Goldstein, 1983; Kendall, 1990; Choudhari and Streett, 1990;
Saric and Reed, 1994; Choudhari, 1994, 1996). A combination of receptivity models with stability ana-
lyses based on the parabolized stability equations (PSE) (Bertolotti and Herbert, 1991; Chang et al., 1991)
makes feasible the amplitude method (Mack, 1975a,b) coupling characteristics of external disturbances with
the transition onset locus. Much progress toward the subsonic and moderate supersonic transition predic-
tion is partially due to the fact that the disturbance spectrum is relatively simple. For example, transition
on subsonic/supersonic cones, at plates, axisymmetric bodies at zero angle of attack, and airfoils is due
to excitation and downstream amplication of TollmienSchlichting waves, which belong to the rst mode
according to the classication of Mack (1969). Other modes are stable and their eigenvalues are essentially
different from the rst mode. In this case, interaction between TollmienSchlichting waves and other modes
can be neglected. Such a single-mode approach is widely used in the e
n
-method (Malik, 1989) and PSE
analyses (Bertolotti and Herbert, 1991; Chang et al., 1991).
In contrast with subsonic and moderate supersonic ows, the initial phase of hypersonic transition reveals
the following new features:
(i) Besides the rst mode, second and higher modes can coexist in the boundary layer. They belong to
the family of trapped acoustic waves. Once the second mode sets in, it becomes the dominant insta-
bility since its growth rate tends to exceed that of the rst mode. For insulated surfaces, this occurs
for Mach numbers M > 4. For cooled surfaces, the second mode can dominate at an even lower Mach
Prehistory of Instability in a Hypersonic Boundary Layer 361
number. The second-mode instability was experimentally observed by Kendall (1975), Demetriades
(1974), Stetson et al. (1983), and Stetson and Kimmel (1992). The Mach 8 stability and transition ex-
periments of Stetson et al. (1983), see also Stetson and Kimmel (1992), for the boundary layer on
a sharp cone showed that the unstable high-frequency second mode plays a major role in the conical
boundary-layer transition. These data are consistent with the stability calculations of Mack (1987) and
Gasperas (1989).
(ii) The analysis of Guschin and Fedorov (1989, 1990) for inviscid disturbances showed that the second-
mode instability is associated with synchronization between the rst and second modes; i.e., their
frequencies and phase speeds coincide at some point. Near the synchronismpoint, the disturbance spec-
trum splits into two branches. In the branch point vicinities, the normal mode decomposition is not valid
and should be replaced by a local solution accounting for interaction between modes due to nonparallel
effects.
(iii) The stability and transition experiments of Stetson et al. (1991) in a planar boundary layer showed
that low-frequency disturbances are growing despite the fact that the linear stability calculations in-
dicated that these rst-mode disturbances should be stable (Gasperas, 1989). Stetson et al. (1991)
reported that there was no evidence of any second-mode harmonics in the planar boundary layer. The
major disturbances are the low-frequency disturbances, which are growing in a frequency band ex-
pected to be stable. These observations are consistent with the at-plate results of Kendall (1975)
obtained at Mach numbers 3, 4.5, and 5.6. Kendall (1975) indicated that uctuations of all frequen-
cies were observed to grow monotonically larger in the region of a boundary layer extending from
the at-plate leading edge to the predicted location of instability; i.e., in a region where no growth
was expected. In this connection, Mack (1975) developed the forcing theory, which was successfully
applied to Kendalls Mach 4.5 planar data. However, Mack noted that the major difculty in the use
of the forcing theory is that forced disturbances are distinct from free disturbances, and the process
by which the former becomes the latter is unknown. This receptivity problem was analyzed by Fe-
dorov and Khokhlov (1991). They found that the rst and second modes are synchronized with acoustic
waves near the at-plate leading edge. Synchronization leads to effective excitation of the boundary-
layer modes by freestream noise through the mechanism, which is qualitatively different from the
subsonic case.
(iv) For moderate supersonic speeds, the dominant boundary-layer disturbances have relatively low fre-
quencies. In conventional wind tunnels, their major source is associated with acoustic waves radiated
by turbulent boundary layers on the nozzle walls. That causes transition Reynolds numbers observed
in conventional supersonic wind tunnels to be signicantly lower than ones measured in free ights.
For hypersonic transition, wind-tunnel experiments present a different situation (Stetson and Kim-
mel, 1992). The second-mode waves have relatively high frequency. Since most of the freestream
acoustic energy is concentrated in a low-frequency band, the noise intensity rapidly decreases with
increasing frequency, and it becomes very small (below the instrumentation noise) in the second-
mode frequency band. Therefore, a noisy hypersonic wind tunnel can be quiet for second-mode
dominated transition. In this case, freestream turbulence (vortical disturbances) and entropy spottiness
(temperature disturbances) may play an important role in boundary-layer receptivity. Under free-ight
conditions, when external acoustic disturbances are negligibly small, atmospheric turbulence can be
also important.
These examples show that extension of subsonic and low supersonic stability concepts and transition
prediction methods to hypersonic speeds is not straightforward. Most of the features listed above relate to
disturbance evolution in regions located upstream from the lower neutral branch. However, previous stud-
ies have been focused on downstream regions where instability sets in. The objective of this paper is to
analyze the instability prehistory in the upstream regions. In Section 2 we formulate a basic formalism
of the multiple-mode method following the papers of Zhigulev et al. (1980), Tumin and Fedorov (1983),
and Zhigulev and Tumin (1987). In contrast with the single-mode approach, the multiple-mode method
allows us to analyze receptivity mechanisms and intermodal exchanges in the synchronism regions. In
Section 3 we discuss the disturbance spectrum topology for hypersonic boundary layers on a sharp cone
or at plate. We identify singular regions associated with synchronization of boundary-layer modes with
acoustic waves (Region 1), and vorticity/entropy waves (Region 2), as well as synchronization between the
362 A.V. Fedorov and A.P. Khokhlov
rst and second modes (Region 3). In Section 4 we study coupling between the rst and second modes and
establish the intermodal exchange rule for Region 3. In Section 5 we analyze the spectrum topology in Re-
gion 2 and discuss receptivity to freestream turbulence and entropy spottiness. In Section 6 we summarize
results.
2. Basic Formalism
For completeness we briey describe the multiple-mode method developed by Zhigulev et al. (1980), Tu-
min and Fedorov (1983), and Zhigulev and Tumin (1987) for compressible boundary layers. We consider
a two-dimensional boundary-layer ow with mean velocity components (U, V ) =(U

, V

)/U

e
, pressure
P = P

/(

e
U
2
e
), and temperature T =T

/T

e
. Due to downstream growth of the boundary-layer thickness,
the mean-ow proles are functions of the longitudinal coordinate x = x

/L

:
U =U(x, y), V =

V(x, y), P = P(x), T = T(x, y) , (2.1)
where y = y

/ is the coordinate normal to the wall surface, =


_

e
L

/U

e
is the boundary-layer scale, and
L

is the streamwise distance from the body leading edge to the lower neutral branch. A two-dimensional
disturbance is expressed in the vector form
Z(x, y, t) =
_
u,
u
y
, v, p, ,

y
,
u
x
,

x
,

x
_
T
, (2.2)
where t =t

e
/ is the time, u and are the streamwise and vertical velocities, p is the pressure, and is
the temperature. For harmonic disturbances of the angular frequency =

/U

e
, we have
Z =A(x, y) exp(it) . (2.3)
The disturbance amplitude A is a solution of the linearized NavierStokes equations, which can be written in
matrix-operator form:
H
0
AH
2
A
x
=0 , (2.4a)
H
0
=

y
L
0

y
+L
1

y
H
1
H
3
. (2.4b)
The 99 matrices L
0
, L
1
, H
1
, H
2
, and H
3
depend on mean-ow characteristics; their explicit form is given
in Zhigulev and Tumin (1987). Elements of the matrix H
3
are proportional to the terms U/x, T/x,

V,
and

V/y associated with nonparallel effects. They are of the order of =1/R
1
, where the Reynolds num-
ber is R
1
=
_
U

e
L

e
. If the disturbance amplitude is identied in a certain cross section, x = x
0
, then the
boundary conditions are
A(x
0
, y) =A
0
(y) , (2.5a)
A
1
(x, 0) = A
3
(x, 0) = A
5
(x, 0) =0 , (2.5b)
|A(x, y)| <, y , (2.5c)
where (2.5b) indicates no-slip conditions and zero temperature perturbation on the wall of high thermal
conductivity. A solution of problem (2.4), (2.5) can be represented as the elementary wave decomposition:
A =

(x)A

(x, y) exp [

(x)],

(x) =i
1
_
x
x
0
(x) dx , (2.6)
where

denotes summation over discrete spectrum and integration over continuous spectrum. The
eigenvectors A

belong to the biorthogonal system {A

(x, y), B

(x, y)} resulting from solutions of the


Prehistory of Instability in a Hypersonic Boundary Layer 363
direct problem

y
_
L
0
A

y
_
+L
1
A

y
= H
1
A

+iH
2
A

, (2.7a)
A
1
(x, 0) = A
3
(x, 0) = A
5
(x, 0) =0 , (2.7b)
|A

(x, y)| <, y , (2.7c)


and the adjoint problem

y
_
L

0
B

y
_
L

1
B

y
= H

1
B

i H

2
B

, (2.8a)
B
2
(x, 0) = B
4
(x, 0) = B
6
(x, 0) =0 , (2.8b)
|B

(x, y)| <, y . (2.8c)


Here an asterisk denotes a conjugate matrix and an overbar denotes complex conjugate quantities. Prob-
lems (2.7) and (2.8) are formulated for a locally parallel ow; i.e., the term H
3
in (2.4b) is neglected. The
eigenvectors satisfy the orthogonality condition
_
H
2
A

, B

_
=

, (2.9)
where the scalar product is dened as
A,B = lim
k0
_

0
exp (y)(A,B) dy, (A,B) =
9

j=1
A
j

B
j
, > 0 . (2.10)
Here

is the Kronecker symbol if one of the eigenvalues , belongs to the discrete spectrum;

=
() is the delta-function if both eigenvalues belong to the continuous spectrum.
Substituting (2.6) into (2.4), multiplying by B

, and accounting for (2.9), we obtain the ordinary differ-


ential equation system for the amplitude coefcients c

(x):
dc

dx
=

(x)W

(x) exp [

] , (2.11)
c

(x
0
) =H
2
A
0
, B

,
where the matrix elements are determined as
W

=
_
H
2
A

x
, B

_
H
3
A

, B

_
. (2.12)
For arbitrary normalization of the eigenvectors A

, B

, these elements are expressed in the form


W

=
_
H
2
(A

/x) , B

_
+
_
H
3
A

, B

_
H
2
A

, B

. (2.13)
The diagonal elements W

describe nonparallel effects on the evolution of a single mode. The nondiago-


nal elements W

are associated with interaction between the modes of eigenvalues and . If one of these
modes belongs to the continuous spectrum, then W

couple external disturbances with the boundary-layer


modes (receptivity problem). If both modes belong to the discrete spectrum, these elements are associated
with the intermodal exchange. If the mean ow is parallel, then all matrix elements W

=0 and no coupling
occurs between modes.
In the framework of this formalism, analyses of receptivity and intermodal exchange include the follow-
ing steps: (1) identify the biorthogonal eigenvector system {A

, B

} for discrete and continuous spectrums


of (x, ); (2) specify the initial conditions at x = x
0
; (3) simplify the system (2.11), neglecting modes
which are not important for a particular problem (cutoff procedure); (4) calculate eigenvalues, eigenfunc-
tions, and the matrix elements (2.13); (5) calculate the amplitude coefcients c

(x). Note that extension of


this approach to three-dimensional disturbances is straightforward (Zhigulev and Tumin, 1987).
364 A.V. Fedorov and A.P. Khokhlov
3. Disturbance Spectrum
Tumin and Fedorov (1983) showed that the local homogeneous problem(2.7) can be reduced to the standard
stability problem:
dz

dy
= G
0
z

, (3.1a)
z
1
= z
3
= z
5
=0, y =0 , (3.1b)
|z

| <, y , (3.1c)
where z

= (A
1
, A
2
, A
3
, A
4
, A
5
, A
6
)
T
. For two-dimensional disturbances, the 66 matrix G
0
de-
pends on mean-ow characteristics, frequency , eigenvalue , and Reynolds number R =
_
U

e
x

e

R
1

x. Its nonzero elements are presented in many papers; see, for example, Mack (1969, 1975a) and
Nayfeh (1980).
Tumin and Fedorov (1983) showed that besides the discrete spectrum there are seven branches of con-
tinuous spectrum: three branches correspond to waves propagating upstream with rapid decay, two branches
correspond to acoustic waves propagating downstream, and two branches correspond to vorticity and entropy
waves propagating downstream. Further discussion addresses disturbances propagating downstream only.
For large Reynolds numbers R, the vorticity,

, and entropy,
e
, wave numbers are approximated as

=+i
k
2
+
2
R
+O
_
R
2
_
, (3.2)

e
=+i
k
2
+
2
RPr
+O
_
R
2
_
, (3.3)
where Pr is the Prandtl number; k is the y-component of the wave number for disturbances outside the
boundary layer, y 1. In this region the eigenvectors of vorticity and entropy waves oscillate as
z
,e
=q
(1)
,e
exp(iky) +q
(2)
,e
exp (iky) , k > 0 ,
where the vectors q
(1)
,e
and q
(2)
,e
are constant or weak functions of x, y. According to (3.2) and (3.3) vor-
ticity and entropy waves propagate downstream with the phase speed c
r
/
r
=1+O(R
2
), and slowly
dissipate due to viscosity.
For acoustic waves in a supersonic ow with M
e
> 1, the wave number
a
is in the ranges
Re(
a1
)
M
e

1+M
e
+O
_
R
1
_
, Re(
a2
)
M
e

1M
e
+O
_
R
1
_
. (3.4)
The rst range corresponds to fast acoustic waves with the phase speed c 1+1/M
e
, and the second range
corresponds to slow acoustic waves with the phase speed c 11/M
e
.
To illustrate basic features of the hypersonic boundary-layer spectrumwe consider self-similar boundary-
layer proles on a sharp cone (Hayes and Probstein, 1959). The conical mean-ow characteristics are
obtained using the Mangler transformation according to which the planar boundary-layer thickness is larger
than the conical boundary-layer thickness by

3. All data presented hereafter can be adjusted to the at-
plate boundary layer using the scaling: R R/

3, /

3. In all calculations, the uid is a perfect


gas of the specic heats ratio = 1.4 and Pr =0.72; gas temperature at the upper boundary-layer edge is
T

e
=70.26 K. The viscosity temperature dependency is approximated by Sutherlands law:
(T) =
(1+S)
(T +S)
T
3/2
,
where S = 110/T

e
; =

e
is the nondimensional viscosity. The second viscosity is assumed to be

v
=0.8. Effects associated with hypersonic viscousinviscid interaction are neglected, since the interac-
tion parameter = M
3
e
/R (see Hayes and Probstein, 1959) is less than 0.2 for the cases discussed hereafter.
The eigenvalues of the rst and second modes as well as the boundaries of the continuous spectrum
are shown in Figure 1 for the Mach number M
e
=5.5, wall temperature ratio T
w
/T
ad
=0.1, and frequency
Prehistory of Instability in a Hypersonic Boundary Layer 365
Figure 1. Disturbance spectrum (R) at F =10
4
, M
e
=5.5, T
w
/T
ad
=0.1.
parameter F =10
4
. Near the leading edge (Region 1 in Figure 1(a)),
1,2
tend to the lower and upper
boundaries of the acoustic spectrum, respectively. Asymptotic analysis of Fedorov and Khokhlov (1991) and
our numerical calculations conducted in a wide range of Mach numbers and wall temperature ratios showed
that this upstream asymptotic behavior is typical for hypersonic boundary layers and can be used for clas-
sication of the boundary-layer modes. We call a discrete mode of the phase speed c
r
1+1/M
e
, x 0,
Mode 1, and a discrete mode of the phase speed c
r
11/M
e
, x 0, Mode 2. Note that this classication
is different from that introduced by Mack (1969). To avoid misunderstanding we call the rst/second mode
of Macks classication the Mack rst/second mode.
Fedorov and Khokhlov (1991) showed that synchronization of Modes 1 and 2 with acoustic waves of the
phase speed c
r
=11/M
e
leads to strong excitation of the boundary-layer instability near the leading edge
(Region 1 in Figure 1(a)). They analyzed an asymptotic structure of this region and derived analytical ex-
pressions for the initial amplitude of Modes 1 and 2. Note that this receptivity mechanism is different from
that investigated by Goldstein (1983) for subsonic boundary layers.
Near the point R 3600 (Region 2 in Figure 1(a)), Mode 1 is synchronized with vorticity and entropy
waves of the phase speed c
r
=1. In this region, freestream vorticity and entropy disturbances may effec-
tively generate Mode 1 and, conversely, Mode 1 may induce vortices and entropy spots. This mechanism is
discussed in Section 5.
In the vicinity of R 3900 (Region 3 in Figure 1(a)), the phase speeds of Modes 1 and 2 are very
close to each other. Due to this synchronization the disturbance spectrum branches out: Mode 1 becomes
unstable (this instability is related to the Mack second mode), whereas Mode 2 becomes more stable. The
spectrum topology may be different from that shown in Figure 1. As an example, Figure 2 illustrates the
eigenvalue behavior for M
e
=5.95, T
w
/T
ad
=0.1, and F =10
4
. In this case, Mode 1 becomes more stable
whereas Mode 2 becomes unstable (again this instability is related to the Mack second mode). The spectrum
topologies and disturbance behavior in the synchronism Region 3 are discussed in the next section.
4. Synchronism of Discrete Modes (Region 3)
A qualitative behavior of the disturbance spectrum shown in Figures 1 and 2 was considered by Guschin
and Fedorov (1989, 1990) in the framework of inviscid stability theory. They assumed that in Region 3 the
dispersion relation can be approximated as
( a
1
)( a
2
) =a
0
, (4.1)
366 A.V. Fedorov and A.P. Khokhlov
Figure 2. Disturbance spectrum (R) at F =10
4
, M
e
=5.95, T
w
/T
ad
=0.1.
where =
0
and =
0
= F(RR
0
) = F

R are the local wave number and angular frequency,
respectively; a
0
, a
1
, and a
2
are constants. Adjusting these constants Guschin and Fedorov (1989) showed
that (4.1) captures basic features of the discrete spectrum behavior in the synchronism region. A similar
dispersion relation was derived by Smith and Brown (1990) using an asymptotic method. Note that (4.1)
arises in many physical problems. For example, it is widely used for stability analyses in plasma physics,
see Fedorchenko and Kotsarenko (1981). The solution of (4.1) has two branches:

1,2
=
a
1
+a
2
2
F

R
_
1
4
(a
1
a
2
)
2
F
2
R
2
+a
0
. (4.2)
The eigenvalues
1,2
(

R) are schematically shown in Figure 3 for the case of a
0
< 0, a
1
> a
2
> 0, and xed
frequency F. The branch points are located on the real axis of the complex

R-plane and are determined as

R
b1,2
=
2

|a
0
|
F(a
1
a
2
)
. (4.3)
If these points are bypassed from the upper side as shown in Figure 3, then Mode 1 is stable, Im(
1
) >0, and
Mode 2 is unstable, Im(
2
) <0, in the range

R
b1
<

R <

R
b2
. This branching is consistent with the numerical
result presented in Figure 2. As is shown below, the branch point R
b1
is slightly shifted from the real axis to
the lower half of the complex R-plane and bypassed from the upper side. Figure 1 illustrates another case
when the branch point R
b1
is shifted to the upper half and bypassed from the lower side.
Further analysis is focused on disturbance behavior near the upstream branch point x
b1
=(R
b1
/R
1
)
2
. In
its vicinity, the eigenvalues of Modes 1 and 2 are approximated as

1,2
=
0

X +. . . , (4.4)
where X = x x
b1
is a local variable, and is constant. Assuming that coupling between these modes
weakly depends on other disturbances, we use the two-mode approximation
A =c
1
(x)A
1
(x, y) exp [
1
(x)] +c
2
(x)A
2
(x, y) exp[
2
(x)] , (4.5)

1,2
=i
1
_
x
x
b1

1,2
(x) dx , (4.6)
Prehistory of Instability in a Hypersonic Boundary Layer 367
Figure 3. Topology of the dispersion relation (4.1) in the synchronism region.
where the eigenfunctions A
1,2
are solutions of the homogeneous problem
(H
0
iH
2
) A
j
=0 , (4.7a)
A
j1
(x, 0) = A
j3
(x, 0) = A
j5
(x, 0) =0 (4.7b)
|A
j
(x, y)| 0, y j =1, 2 . (4.7c)
The amplitude coefcients c
1,2
satisfy the ordinary differential equation system
dc
1
dx
=c
1
W
11
+c
2
W
12
exp(
2

1
) , (4.8a)
dc
2
dx
=c
2
W
22
+c
1
W
21
exp(
1

2
) . (4.8b)
In the branch point vicinity, the operators H
0
, H
2
and eigenfunctions A
1,2
can be expanded as
H
0
= H
00
+XH
01
+. . . , (4.9a)
H
2
= H
20
+XH
21
+. . . , (4.9b)
A
j
=A
(0)
+(
j

0
) A
(1)
+. . . , j =1, 2 . (4.9c)
Substituting (4.4) and (4.9) into (4.7) we obtain the following problems:
(H
00
i
0
H
20
) A
(0)
=0 , (4.10a)
A
(0)
1
(x, 0) = A
(0)
3
(x, 0) = A
(0)
5
(x, 0) =0 , (4.10b)
|A
(0)
(x, y)| 0 , y ; (4.10c)
(H
00
i
0
H
20
) A
(1)
=iH
20
A
(0)
, (4.11a)
A
(1)
1
(x, 0) = A
(1)
3
(x.0) = A
(1)
5
(x, 0) =0 , (4.11b)
|A
(1)
(x, y)| 0 , y ; (4.11c)

2
(H
00
i
0
H
20
) A
(2)
=i
2
H
20
A
(1)
(H
01
i
0
H
21
) A
(0)
, (4.12a)
A
(2)
1
(x, 0) = A
(2)
3
(x, 0) = A
(2)
5
(x, 0) =0 , (4.12b)
|A
(2)
(x, y)| 0 , y . (4.12c)
368 A.V. Fedorov and A.P. Khokhlov
For nontrivial solutions of problems (4.11) and (4.12), the right-hand sides of (4.11a) and (4.12a) are orth-
ogonal to the adjoint-problemsolution B
(0)
:
_
H
20
A
(0)
, B
(0)
_
=0 , (4.13)
i
2
_
H
20
A
(1)
, B
(0)
_
=
_
(H
01
i
0
H
21
) A
(0)
, B
(0)
_
. (4.14)
Equation (4.13) can be used for searching the branch point, and the constant can be determined
from (4.14).
Similar expansions are made for the adjoint-problem solution B. Substituting them into (2.13) and eval-
uating the scalar products we can express the matrix elements in the form
W
jk
=
(1)
j+k1
4X
+. . . , X 0 ; j =1, 2 , k =1, 2 . (4.15)
In the inner region with the variable =
2/3
X = O(1), system (4.8) is expressed as
dc
1
d
=
1
4
_
c
1
+c
2
exp (2)
_
, (4.16a)
dc
2
d
=
1
4
_
c
1
exp (2) c
2
_
, (4.16b)
() =
2
3
i
3/2
. (4.16c)
Note that (4.16a) and (4.16b) do not depend on the mean-ow proles. Their solution can be used for mod-
eling intermodal exchanges in weakly nonparallel shear layers with the discrete spectrum behavior being
similar to that given by (4.4). Using the transformation
f() =c
1
() exp () +c
2
() exp () , (4.17)
we can reduce system (4.16) to the Airy equation
d
2
f
d
2
+
2
f =0 . (4.18)
For , an asymptotic behavior of f() depends on orientations of Stokes lines l
j
, j =1, 2, 3, which
are determined from the equation Re[()] =0. We consider the Stokes lines conguration schematically
shown in Figure 4. In this case, Modes 1 and 2 are almost neutral in the upstream region < 0. For > 0,
one of them is unstable and another is stable. Numerical examples of this spectrum branching are shown in
Figures 1 and 2.
Using the method developed by Fedoruk (1983) we introduce the canonical domains D
j
: l
j
D
j
, l
k=j

D
j
and specify the branches of () using the condition: Im() > 0 at l
j
D
j
. Then the real part Re
is positive to the right of the Stokes line l
j
, and negative to the left of l
j
. Then the solution f() of (4.18) has
the asymptotic form
f S
j

1/4
[a
j
exp () +b
j
exp ()], || , (4.19)
where D
j
. The constants S
j
satisfy the conditions |S
j
| =1 and arg(S
j

1/4
) =0 at l
j
, j =1, 2, 3. The
constants a
j
and b
j
are coupled as
_
a
1
b
1
_
=exp
_
i

6
_
_
0 1
1 i
_ _
a
3
b
3
_
, (4.20a)
_
a
j+1
b
j+1
_
=exp
_
i

6
_
_
0 1
1 i
__
a
j
b
j
_
, j =1, 2 . (4.20b)
Prehistory of Instability in a Hypersonic Boundary Layer 369
Figure 4. Stokes lines l
j
and canonical domains D
j
in the branch point vicinity.
Figure 5. Intermodal exchange diagrams: (a) stable and (b) unstable mode translation.
Equations (4.20a) and (4.20b) determine the intermodal exchange rule for Modes 1 and 2 passing through the
interaction region |x x
b1
| = O(
2/3
) and/or crossing one of the Stokes lines. In the rst-order approxima-
tion, this rule is valid for any weakly nonparallel mean ow, which has a similar local topology of its discrete
spectrum. It can be expressed in the generic form
_
c
+
st
c
+
un
_
=exp
_
i

6
_

_
0 1
1 i
_

_
c

st
c

un
_
, (4.21)
where (c

st
, c

un
) and (c
+
st
, c
+
un
) are amplitudes of stable and unstable modes upstream and downstream from
the interaction region, respectively. For the case shown in Figure 1, c

st
, c
+
st
correspond to Mode 2, and c

un
,
c
+
un
correspond to Mode 1. For the case shown in Figure 2, c

st
, c
+
st
correspond to Mode 1, and c

un
, c
+
un
cor-
respond to Mode 2. The intermodal exchange diagram shown in Figure 5 indicates that, in the rst-order
approximation, a stable mode excites an unstable mode of the amplitude |c
+
un
| |c

st
|. In turn, an unstable
mode excites a stable mode of the amplitude |c
+
st
| |c

un
| and departures from the exchange region with the
amplitude |c
+
un
| |c

un
|. In both cases the unstable mode occurs in the downstream region. In the rst-order
approximation, the intermodal exchange rule (4.21) does not depend on the parameter associated with non-
parallel effects. However, the local amplitude grows proportionally to
1/6
and the interaction region length
increases proportionally to
2/3
as 0.
If the interaction between Modes 1 and 2 is neglected, then (4.16a) and (4.16b) are decoupled:
dc
j
d
=
1
4
c
j
, j =1, 2 . (4.22)
The amplitude coefcients tend to innity, c
j

1/4
, as 0. This singularity indicates that the
single-mode approximation (Nayfeh, 1980), which is widely used for nonparallel stability calculations, is
not valid in the branch-point vicinity. The two-mode approximation is required to resolve this singularity and
obtain a uniformly valid solution.
370 A.V. Fedorov and A.P. Khokhlov
Figure 6. The branch point R
b1
as a function of frequency parameter F; M
e
=6.8, T
w
= T
ad
.
To capture the branch points, stability calculations should be extended to complex Reynolds numbers.
For self-similar boundary-layer ows, this extension is straightforward, because the stability matrix G
0
of
(3.1a) depends on R in a parametrically explicit form only. In this case the eigenvalue problem (3.1) can
be solved at complex Reynolds numbers with the help of any numerical algorithm appropriate for stability
calculations. The branch Reynolds number R
b1
is calculated from (4.13) using the Newton iteration tech-
nique. Distributions Re[R
b1
(F)] and Im[R
b1
(F)] are shown in Figure 6 for the case of local Mach number
M
e
=6.8 and wall temperature ratio T
w
/T
ad
=1. Since the branch point is in the lower half of the complex
R-plane, the spectrum topology is similar to the case shown in Figure 2.
Figure 7 shows the branch Reynolds number R
b1
as a function of the local Mach number M
e
for various
wall temperature ratios T
w
/T
ad
at the disturbance frequency parameter F =10
4
. For the cooled wall case,
T
w
/T
ad
=0.1, the imaginary part Im(R
b1
) changes its sign at M
e
5.9. In the region M
e
> M
e
, the branch
point is in the lower half of complex R, and the spectrumtopology corresponds to the case shown in Figure 2.
In the region M
e
< M
e
, the branch point is in the upper half, and the spectrum topology corresponds to the
case shown in Figure 1. An innitesimal variation of M
e
(or any other mean-ow parameter) near its critical
Figure 7. The branch point R
b1
as a function of local Mach number M
e
at various wall temperature ratios T
w
/T
ad
; F =10
4
.
Prehistory of Instability in a Hypersonic Boundary Layer 371
value M
e
causes a jump from one topology to another. However, the intermodal exchange rule (4.21) (see
also Figure 5) indicates that the unstable mode amplitude is approximately the same in both cases; i.e., the
Mack second-mode amplitude is weakly sensitive to this jump.
If the real x-axis crosses the local region |x x
b1
| = O(
2/3
), i.e., | Im(x
b1
)| = O(
2/3
) and | Im(R
b1
)| =
O(R
b1
1/3
), the disturbance is strongly affected by the branch point. Its group velocity V
g
=( Re()/)
1
tends to minus or plus innity and then jumps to a certain positive value as the Reynolds number crosses the
branch point. This singularity can cause difculties in stability analyses. For example, the temporal growth
rate
i
cannot be converted to the spatial growth rate using the Gaster transformation =
i
/|Re V
g
|.
PSE algorithms may also be affected because the disturbance characteristics rapidly vary near the branch
point. For example, in the PSE approach of Chang et al. (1991), the disturbance in the vicinity of a given
location x
1
is expressed as
(x, y, t) =(x, y) exp
_
i
__
x
x
1

1
dx t
__
. (4.23)
Hereafter in this section we use the notations of Chang et al. (1991). The shape function is approximated
by the Taylor series expansion
(x, y) =(x
1
, y) +

x
(x
1
, y)(x x
1
) +. . . . (4.24)
In order to make the marching scheme well-posed, the wave number is updated as
=
1
i
1
(x
1
, y
m
)

x
(x
1
, y
m
) , (4.25)
where y
m
corresponds to the maximum amplitude of streamwise velocity. If the location x
1
coincides with
the branch point x
b
, then the expansion (4.24) is not valid. Moreover, the wave-number correction given by
(4.25) tends to innity as (/x)/ (x
1
x
b
)
1/4
.
5. Synchronism of Mode 1 with Entropy/Vorticity Disturbances (Region 2)
In Region 2 depicted in Figures 1(a) and 2(a), the eigenvalue
1
crosses the ray c
r
= 1 relevant to
vorticity/entropy waves. This region is shown on an enlarged scale in Figure 8 for the Mach number
M
e
=6.8, wall temperature ratio T
w
/T
ad
=1, and disturbance frequency F =10
4
. Figure 9 schematically
illustrates the spectrum topology in the complex -plane. Since the entropy-wave branch merges with the
vorticity-wave branch as R , both branches can be approximated as one branch cut. The eigenvalue
1
enters to the left side of this cut at R = R

0 and departs from its right side at R = R

+0, where R

1515
can be treated as a synchronism point. The imaginary part
1i
has the jump (
1i
) 0.3410
2
, while
the real part remains smooth:
1r
(R

0) =
1r
(R

+0). The eigenfunctions of the x-component velocity,


u() =Re[z
1
()], where = y/

x is the similarity variable, are shown in Figure 10 for various Reynolds


numbers in the vicinity of R

. These functions are normalized by the condition max| u()| =1. It is seen
that velocity oscillations penetrate outside the boundary layer as R R

, that indicates synchronization of


Mode 1 with entropy/vorticity waves. It is also seen that the eigenfunction at R = R

0 (Line 3) is different
from that at R = R

+0 (Line 4).
Nonparallel effects can cause a strong interaction between Mode 1 and vorticity/entropy waves in
Region 2. Due to the presence of a continuous spectrum it is difcult to simplify and solve (2.11). However,
rough estimates of this interaction can be made as follows. We introduce a virtual upper boundary, =
s
,
located in the inviscid outer-ow region, and impose the homogeneous conditions
z
1
= z
3
= z
5
=0, =
s
. (5.1)
Then the continuous-spectrum branches translate to subsets of the discrete spectrum.
Numerical analysis of the problem (3.1a), (3.1b), and (5.1) shows that the mode with eigenvalue
m
,
which belongs to the vorticity/entropy wave subset and satises the synchronism condition,
m
(R

)
372 A.V. Fedorov and A.P. Khokhlov
Figure 8. Disturbance spectrum (R) in the vicinity of the synchronism point R

1515; M
e
=6.8, T
w
= T
ad
.
Figure 9. Spectrum topology near the synchronism point R

.
Figure 10. Eigenfunctions u() of Mode 1; 1, R =1369; 2, R =1438; 3, R =1507; 4, R =1542; 5, R =1611; 6, R =1681; the
synchronism point R

1515.
Prehistory of Instability in a Hypersonic Boundary Layer 373
Figure 11. Mode transformations in the synchronism region; the virtual upper boundary y

s
=4

; dashed line, vorticity/entropy


waves for y
s
.
Figure 12. Schematic picture of vorticity/entropy disturbance swallowing in synchronism regions near the points R

(F
1
) and R

(F
2
),
F
1
> F
2
.

1
(R

+0) as y
s
, translates to Mode 1 in the region R > R

. In a similar way, Mode 1 with the eigen-


value
1
in the upstream region R < R

, translates to the vorticity/entopy wave with the eigenvalue


n
,
which satises the synchronism condition
n
(R

)
1
(R

0) as y
s
. These translations are shown
in Figure 11 for the virtual boundary y

s
=4

, where

is the boundary-layer displacement thickness. In


this case the region between the upper boundary-layer edge and the virtual boundary contains about 12
wavelengths of vorticity/entropy oscillations in the y-direction. In the upstream region R < R

1515,
the normal mode eigenvalue
m
coincides with the vorticity/entropy wave eigenvalue (dashed line), which
satises the synchronism condition

(R

, k

) =
1
(R

+0) and/or
e
(R

, k
e
) =
1
(R

+0) . (5.2)
These ndings lead to the assumption that vorticity/entropy waves can effectively generate Mode 1 in the
vicinity of R

. Further downstream (in Region 3), Mode 1 excites Mode 2 due to the intermodal exchange
discussed in Section 4. One of these modes is the Mack second mode. Thus, vorticity (entropy) waves with
the wave number

(
e
), in the vicinity of the synchronism values (5.2), can be swallowed by the bound-
ary layer in Region 2 and translated to the Mack second mode as schematically shown in Figure 12. This
receptivity mechanism occurs near the lower neutral branch of the Mack second mode and provides ini-
tial amplitudes for unstable waves just at the beginning of their amplication. In order to determine the
wave-number range involved in the excitation of Mode 1 and to evaluate the receptivity coefcients, further
analysis is needed.
374 A.V. Fedorov and A.P. Khokhlov
6. Discussion
The Transition process is an initial boundary-value problem, which requires initial data for unstable nor-
mal modes. These data are determined from the instability prehistory analysis that allows us to identify
external disturbances, determine the most receptive regions, and predict the normal-mode propagation from
these regions to the instability growth onset. All these phases for the hypersonic boundary layer are different
from those observed at subsonic and moderate supersonic speeds. Hypersonic instability prehistory reveals
the following new features: (1) in the leading edge region, Modes 1 and 2 are synchronized with fast and
slow acoustic waves of the phase speed c
r
=1+1/M
e
and c
r
=11/M
e
, respectively; (2) further down-
stream, Mode 1 is synchronized with the external entropy/vorticity waves of the phase speed c
r
=1; (3)
near the second-mode neutral branch, Mode 1 is synchronized with Mode 2. Frequencies and phase speeds
of synchronized disturbances are equal or very close to each other, that leads to strong interactions between
disturbances due to nonparallel effects.
In Region (1), acoustic waves can effectively excite Modes 1 and 2 (Fedorov and Khokhlov, 1991). This
receptivity mechanism is substantially different from the subsonic case. The theoretical prediction of Fe-
dorov and Khokhlov (1991) is in qualitative agreement with the leading-edge receptivity experiments of
Maslov et al. (1998) conducted on a at plate at Mach 6. Further theoretical and experimental studies are
needed to perform quantitative comparisons.
In Region (2), vorticity and entropy waves, which are synchronized with Mode 1, may be swallowed by
the boundary layer and effectively generate Mode 1. This receptivity mechanism needs further theoretical
studies and experimental verication. We believe that it can compete with receptivity to acoustic distur-
bances in the case when freestream noise is small compared with turbulence and temperature spottiness
in a high-frequency band relevant to transition. Low levels of freestream noise are typical for free ights.
Hypersonic wind tunnels may also be quiet at high frequencies.
In Region (3), Mode 1 is synchronized with Mode 2. The discrete spectrum has branch points located in
the complex R-plane near the real axis. Downstream from the rst branch point R
b1
, Mode 1 or Mode 2 be-
comes unstable depending on the sign of Im(R
b1
). In both cases the unstable mode is equivalent to the Mack
second mode. Near the branch points, the normal mode decomposition is not valid and should be replaced
by a local solution, which provides coupling between input and output amplitudes of the interacting modes.
This coupling is governed by the intermodal exchange rule (4.21), which, in the rst-order approximation,
does not depend on the mean-ow proles and can be applied for various shear layers with similar local top-
ology of their discrete spectrum. Our parametric calculations show that the branch point R
b1
can cross the
real axis. That leads to different branching of the discrete spectrum. Innitesimal variations of a mean-ow
parameter (say the local Mach number M
e
) near its critical value cause a jump from one topology to another.
Moreover, near the branch point the disturbance group velocity and amplitude have a nonmonotonic singular
behavior. These features, along with the spectrum singularity in Region (2), can cause signicant difculties
in stability analyses, which ignore interactions between different modes. The multiple-mode method helps
to resolve these difculties and obtain uniformly valid solutions.
Acknowledgment
Useful discussions with Drs. Roger Kimmel and Jonathan Poggie are gratefully acknowledged. We appreci-
ate the support provided by Drs. Mark Maurice and Charbel Raffoul.
References
Bertolotti, F.P., and Herbert, Th. (1991). Analysis of the linear stability of compressible boundary layers using the PSE. Theoret.
Comput. Fluid Dynamics, 3, 117124.
Chang, C.-L., Malik, M.R., Erlebacher, G., and Hussaini, M.Y. (1991). Compressible Stability of Growing Boundary Layers Using
Parabolized Stability Equations. AIAA Paper No. 91-1636.
Choudhari, M. (1994). Theoretical Prediction of Boundary-Layer Receptivity. AIAA Paper No. 94-2223.
Choudhari, M. (1996). Boundary-Layer Receptivity to Three-Dimensional Unsteady Vortical Disturbances in Free Stream. AIAA
Paper No. 96-0181.
Prehistory of Instability in a Hypersonic Boundary Layer 375
Choudhari, M., and Streett, C. (1990). Boundary Layer Receptivity Phenomena in Three-Dimensional and High-Speed Boundary
Layers. AIAA Paper No. 90-5258.
Demetriades, A. (1974). Hypersonic Viscous Flow over a Slender Cone, Part III: Laminar Instability and Transition. AIAA Paper No.
74-535.
Fedorov, A.V., and Khokhlov, A.P. (1991). Excitation of unstable modes in a supersonic boundary layer by acoustic waves. Izv. Akad.
Nauk SSSR, Mekh. Zhidk. Gaza, No. 4, 6774.
Fedorchenko, A.M., and Kotsarenko, N.Ya. (1981) Absolute and Convective Instability in Plasma and Solids. Moscow, Nauka (in
Russian).
Fedoruk, M.V. (1983). Asymptotic Methods for Ordinary Differential Equations. Nauka, Moscow (in Russian).
Gaponov, S.A., and Maslov, A.A. (1980) Development of Disturbances in Compressible Flows. Nauka, Novosibirsk (in Russian).
Gasperas, G. (1989). Effect of Wall Temperature Distribution on the Stability of Compressible Boundary Layer. AIAA Paper No.
89-1894.
Goldstein, M.E. (1983). The evolution of TollmienSchlichting waves near a leading edge. J. Fluid Mech., 127, 5981.
Guschin, V.R., and Fedorov, A.V. (1989). Qualitative properties of the instability of currents at a wall in the presence of a ow at
high supersonic speeds. In Modeli Mekhaniki Neodnorod. Sisitem, pp. 93116. ITPM SO AN SSSR, Novosibirsk (in Russian).
Guschin, V.R., and Fedorov, A.V. (1990). Excitation and development of unstable disturbances in supersonic boundary layer. Izv.
Akad. Nauk SSSR, Mekh. Zhidk. Gaza, No. 3, 2129 (in Russian).
Hayes, W.D., and Probstein, R.F. (1959). Hypersonic Flow Theory. Academic Press, New York.
Kachanov, Yu.S., Kozlov, V.V., and Levchenko, V.Ya. (1982). Occurrence of Turbulence in Boundary Layer. Nauka, Novosibirsk (in
Russian).
Kendall, J.M. (1975). Wind tunnel experiments relating to supersonic and hypersonic boundary layer transition. AIAA J., 13(3),
290299.
Kendall, J.M. (1990). Boundary Layer Receptivity to Freestream Turbulence. AIAA Paper No. 90-1504.
Kosinov, A.D., Maslov, A.A., and Shevelkov S.G. (1990). Experiments on the stability of supersonic laminar boundary layers.
J. Fluid Mech., 219, 621633.
Mack, L.M. (1969). Boundary-Layer Stability Theory. Part B. Doc. 900-277, JPL, Pasadena, California, May.
Mack, L.M. (1975a). A numerical method for the prediction of high speed boundary-layer transition using linear theory. In NASA
SP-347, pp. 101123.
Mack, L.M. (1975b). Linear stability theory and the problem of supersonic boundary-layer transition. AIAA J., 13(3), 278289.
Mack, L.M. (1984). Boundary Layer Linear Stability Theory. AGARD Report No. 709, June.
Mack, L.M. (1987). Stability of Axisymmetric Boundary Layers on Sharp Cones at Hypersonic Mach Numbers. AIAA Paper
No. 87-1413.
Malik, M.B. (1989). Prediction and control of transition in supersonic and hypersonic boundary layers. AIAA J., 27(11), 14871493.
Malik, M., Zang, T., and Bushnell, D. (1990). Boundary Layer Transition in Hypersonic Flows. AIAA Paper No. 90-5232.
Maslov, A.A., Shiplyuk, A.A., Sidorenko, A.A., and Arnal, D. (1998) Leading Edge Receptivity of Hypersonic Boundary Layer.
Preprint No. 1-97, ITAM of Russian Academy of Sciences Siberian Branch and Center dEtudes et de Recherches de Toulouse,
Novosibirsk.
Morkovin, M.V. (1969). Critical Evaluation of Transition from Laminar to Turbulent Shear Layers with Emphasis on Hypersonically
Traveling Bodies. AFFDL-TTR 68-149, Air Force Flight Dynamics Laboratory, Wright Patterson Air Force Base, Dayton, Ohio.
Nayfeh, A.H. (1980). Stability of three-dimensional boundary layer. AIAA J., 18(4), 406416.
Reshotko, E. (1976). Boundary layer stability and transition. Ann. Rev. Fluid Mech., 8, 311349.
Reshotko, E. (1994). Boundary Layer Instability, Transition and Control. AIAA Paper No. 94-0001.
Saric, W.S., and Reed, H.L. (1994). Leading Edge Receptivity to Sound: Experiments, DNS, and Theory. AIAA Paper No. 94-2222.
Smith, F.T., and Brown, S.N. (1990). The inviscid instability of a Blasius boundary layer at large values of the Mach number. J. Fluid
Mech., 219, 499518.
Stetson, K.F., and Kimmel, R.L. (1992). The Hypersonic Boundary-Layer Stability. AIAA Paper No. 92-0737, January.
Stetson, K.F., Thompson, E.R., Donaldson, J.C., and Siler, L.G. (1983). Laminar Boundary Layer Stability Experiments on a Cone
at Mach 8. Part 1: Sharp Cone. AIAA Paper No. 83-1761.
Stetson, K.F., Kimmel, R.L., Thompson, E.R., Donaldson, J.C., and Siler, L.G. (1991). A Comparison of Planar and Conical
Boundary Layer Stability and Transition at a Mach Number of 8. AIAA Paper No. 91-1639.
Tumin, A.M., and Fedorov, A.V. (1983). Spatial growth of disturbances in a compressible boundary layer. J. Appl. Mech. Tech. Phys.,
24, 548.
Zhigulev, V.N., and Tumin, A.M. (1987). Origin of Turbulence. Nauka, Novosibirsk (in Russian).
Zhigulev, V.N., Sidorenko, N.V., and Tumin, A.M. (1980). Generation of instability waves in a boundary layer by external turbulence.
J. Appl. Mech. Tech. Phys., 21, 774.

You might also like